Uploaded by Tony Wu

Hodgkin Lymphoma - A Comprehensive Overview (2020)

advertisement
Hematologic Malignancies
Series Editor: Martin Dreyling
Andreas Engert
Anas Younes Editors
Hodgkin
Lymphoma
A Comprehensive Overview
Third Edition
Hematologic Malignancies
Series Editor
Martin Dreyling
München, Germany
Andreas Engert • Anas Younes
Editors
Hodgkin Lymphoma
A Comprehensive Overview
Third Edition
Editors
Andreas Engert
Department of Internal Medicine
University Hospital Cologne
Köln, Nordrhein-Westfalen
Germany
Anas Younes
Lymphoma Service
Memorial Sloan Kettering Cancer
Center
NY, USA
ISSN 2197-9766 ISSN 2197-9774 (electronic)
Hematologic Malignancies
ISBN 978-3-030-32481-0 ISBN 978-3-030-32482-7 (eBook)
https://doi.org/10.1007/978-3-030-32482-7
© Springer Nature Switzerland AG 2020
This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or
part of the material is concerned, specifically the rights of translation, reprinting, reuse of
illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way,
and transmission or information storage and retrieval, electronic adaptation, computer software,
or by similar or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are
exempt from the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in
this book are believed to be true and accurate at the date of publication. Neither the publisher nor
the authors or the editors give a warranty, expressed or implied, with respect to the material
contained herein or for any errors or omissions that may have been made. The publisher remains
neutral with regard to jurisdictional claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface
Hodgkin lymphoma has become one of the best curable malignancies in both
adult and pediatric patients. More than 80% of all patients can be cured with
risk-adapted treatment including chemotherapy and radiotherapy. This progress is largely due to the development of multi-agent chemotherapy and the
improvements in radiotherapy.
Due to the high cure rate and the young age of most patients affected,
Hodgkin lymphoma has also become a model for studying long-term toxicity
of radiotherapy and chemotherapy that may impact quality of life and/or survival. Future treatment should continue to balance the need for improving the
cure rate while reducing treatment-related side effects. In addition, there are
a number of relevant physical and psychosocial issues including infertility
and fatigue that need to be further exploited.
Monoclonal antibodies against this antigen were successfully used for
diagnostic immunophenotyping and exploited therapeutically. This strategy
has come full circle with the advent of the anti-CD30 antibody drug conjugate Brentuximab Vedotin. This drug has shown remarkable responses in
relapsed and refractory classical Hodgkin lymphoma and is also being used
in combination with chemotherapy.
The development of targeted treatment for patients with Hodgkin lymphoma is rapidly evolving. The recent approval of checkpoint inhibitors has
added an additional treatment option for patients with relapsed cHL, opening
the door for new developments including new combinations of chemotherapy
or radiotherapy being evaluated in first line.
This book should give you a comprehensive overview of the most relevant
biology, diagnostic and clinical aspects of Hodgkin lymphoma. We would
like to express our sincere gratitude to all those who have contributed to this
project.
Cologne, Germany
New York, NY, USA Andreas Engert
Anas Younes
v
Contents
Part I Epidemiology and Pathogenesis
1Epidemiology of Hodgkin Lymphoma ���������������������������������������������� 3
Henrik Hjalgrim and Ruth F. Jarrett
2The Role of Viruses in the Genesis of Hodgkin Lymphoma���������� 25
Ruth F. Jarrett, Henrik Hjalgrim, and Paul G. Murray
3Pathology and Molecular Pathology of Hodgkin Lymphoma�������� 47
Andreas Rosenwald and Ralf Küppers
4Microenvironment, Cross-Talk, and Immune Escape
Mechanisms�������������������������������������������������������������������������������������� 69
Lydia Visser, Johanna Veldman, Sibrand Poppema,
Anke van den Berg, and Arjan Diepstra
5What Have We Learnt from Genomics and Transcriptomics
in Classic Hodgkin Lymphoma������������������������������������������������������ 87
Davide Rossi and Christian Steidl
Part II Diagnosis and First-Line Treatment
6Clinical Evaluation�������������������������������������������������������������������������� 99
James O. Armitage and Jonathan W. Friedberg
7Functional Imaging in Hodgkin Lymphoma �������������������������������� 113
Andrea Gallamini, Bruce Cheson, and Martin Hutchings
8Prognostic Factors���������������������������������������������������������������������������� 145
Paul J. Bröckelmann and Lena Specht
9Principles of Radiation Therapy for Hodgkin Lymphoma���������� 171
Joachim Yahalom, Bradford S. Hoppe, Joanna C. Yang,
and Richard T. Hoppe
10Principles of Chemotherapy in Hodgkin Lymphoma������������������ 199
David Straus and Mark Hertzberg
11Treatment of Early Favorable Hodgkin Lymphoma�������������������� 219
Wouter Plattel and Pieternella Lugtenburg
vii
viii
12Treatment of Early Unfavorable Hodgkin Lymphoma���������������� 237
Marc P. E. André and Andreas Engert
13Treatment of Advanced-Stage Hodgkin Lymphoma�������������������� 249
Alexander Fosså, René-Olivier Casasnovas,
and Peter W. M. Johnson
14Optimizing Decision Making in Hodgkin Lymphoma������������������ 265
Susan K. Parsons, Joshua T. Cohen, and Andrew M. Evens
Part III Special Clinical Situations
15Pediatric Hodgkin Lymphoma�������������������������������������������������������� 277
Georgina W. Hall, Cindy Schwartz, Stephen Daw,
and Louis S. Constine
16The Management of Older Patients with Hodgkin
Lymphoma���������������������������������������������������������������������������������������� 297
Boris Böll and Andrew M. Evens
17Nodular Lymphocyte-Predominant Hodgkin Lymphoma ���������� 317
Dennis A. Eichenauer and Ranjana H. Advani
18The Management of Hodgkin Lymphoma
During Pregnancy���������������������������������������������������������������������������� 325
Veronika Bachanova and Joseph M. Connors
19The Management of HIV-Hodgkin Lymphoma���������������������������� 335
Marcus Hentrich and Michele Spina
Part IV Relapsed and Refractory Disease
20Relapsed and Refractory Hodgkin Lymphoma���������������������������� 351
Bastian von Tresckow and Craig Moskowitz
21Allogeneic Transplantation for Relapsed Hodgkin
Lymphoma���������������������������������������������������������������������������������������� 365
Anna Sureda, Martina Pennisi, and Paolo Corradini
22Targeting CD30 in Patients with Hodgkin Lymphoma���������������� 381
Anita Kumar, Stefano Pileri, Anas Younes,
and Andreas Engert
23Hodgkin Lymphoma and PD-1 Blockade�������������������������������������� 395
Reid Merryman, Philippe Armand, and Stephen Ansell
24Other New Agents for Hodgkin Lymphoma���������������������������������� 411
Alison J. Moskowitz and Anas Younes
Part V Survivorship
25Quality of Life in Hodgkin Lymphoma������������������������������������������ 419
Stefanie Kreissl, Hans-Henning Flechtner,
and Peter Borchmann
Contents
Contents
ix
26Second Malignancy Risk After Treatment of Hodgkin
Lymphoma���������������������������������������������������������������������������������������� 429
Michael Schaapveld, David C. Hodgson,
and Flora E. van Leeuwen
27Cardiovascular and Pulmonary Late Effects�������������������������������� 465
Berthe M. P. Aleman and David J. Cutter
28Gonadal Dysfunction and Fertility Preservation
in Hodgkin Lymphoma Patients ���������������������������������������������������� 485
Karolin Behringer and Michael von Wolff
29Cancer-Related Fatigue in Hodgkin Lymphoma�������������������������� 501
Stefanie Kreissl, Anton Hagenbeek, Hans Knoop,
and Peter Borchmann
Part I
Epidemiology and Pathogenesis
1
Epidemiology of Hodgkin
Lymphoma
Henrik Hjalgrim and Ruth F. Jarrett
Contents
1.1
Introduction 4
1.2
Definition and Histological Classification (WHO) 4
1.3
1.3.1
1.3.2
1.3.2.1
1.3.2.2
1.3.2.3
1.3.3
1.3.4
1.3.5
Hodgkin Lymphoma Occurrence Overall Incidence Age-Specific Incidence Patterns Vary Geographically Historical Patterns Modern Age-Specific Incidence Patterns Age-Specific Incidence Patterns for Hodgkin Lymphoma Subtypes Incidence Trends Classifications for Epidemiological Studies: Multi-disease Models Classifications by Age at Diagnosis, Histology and Tumour
Epstein-Barr Virus Status Overlap Between Epidemiological Classifications of Hodgkin
Lymphoma 1.3.6
amilial Accumulation of Hodgkin Lymphoma:
F
Genetic Predisposition 1.4.1
Genetic Studies: Genome-­Wide Association Studies 1.4.1.1 Hodgkin Lymphoma Subtype-­Specific Associations
in Genetic Analyses 5
5
5
5
6
7
7
10
10
11
1.4
1.5
isk Factors R
1.5.1
revailing Hypotheses in Hodgkin Lymphoma Epidemiology P
1.5.1.1 Childhood Socio-Economic Environment 1.5.2
Anthropometry 12
12
12
13
13
13
14
H. Hjalgrim (*)
Department of Epidemiology Research, Statens
Serum Institut, Copenhagen, Denmark
Department of Haematology, Copenhagen University
Hospital Rigshospitalet, Copenhagen, Denmark
e-mail: HHJ@ssi.dk
R. F. Jarrett
MRC-University of Glasgow Centre for Virus
Research, Glasgow, UK
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_1
3
H. Hjalgrim and R. F. Jarrett
4
1.5.3
1.5.3.1
1.5.3.2
1.5.3.3
1.5.3.4
1.5.4
1.5.4.1
1.5.4.2
1.5.4.3
Medical History Infections Primary and Secondary Immune Deficiencies Autoimmune and Allergic Disorders Medications Environmental Exposures Ultraviolet Light Tobacco Alcohol 15
15
16
17
17
18
18
18
18
1.6
Conclusion 19
References 1.1
Introduction
The treatment of Hodgkin lymphoma has
become one of the great successes in modern
oncology with cure rates exceeding 90% for
patients with early-stage disease and approaching the same level for patients with advanced
disease [1, 2] (see also elsewhere in this book).
Although Hodgkin lymphoma was among
the first haematological malignancies described
in the literature in 1832 [3], the advances made
in the understanding of the natural history of
Hodgkin lymphoma and its causes are less
impressive.
It might be ventured that the impressive treatment results leave this to be of little consequence;
however, efforts in this regard are continuously
worthwhile. Accordingly, the good prognosis for
Hodgkin lymphoma is entirely dependent on
modern care which is inaccessible in many parts
of the world, where the disease still carries considerable mortality. Moreover, as highlighted by
a growing literature, the high cure rates have
been achieved at the cost of a high frequency of
late and often severe adverse treatment effects
among Hodgkin lymphoma survivors [4].
Consequently, if better understanding of Hodgkin
lymphoma aetiology could help identify means
to its prevention, e.g. through vaccination [5],
much could be gained.
The present chapter gives an overview of the
epidemiology of Hodgkin lymphoma and summarizes current understanding of its risk factors. For
a more detailed review, please be referred to [6].
19
1.2
Definition and Histological
Classification (WHO)
Hodgkin lymphoma is a malignancy which in the
vast majority of cases is derived from germinal
centre B-lymphocytes, with odd cases (possibly)
being of T-cell origin [7].
The current WHO classification recognizes
two main variants of Hodgkin lymphoma, specifically classic Hodgkin lymphoma and nodular
lymphocytic predominant Hodgkin lymphoma
[8, 9].
The two variants of Hodgkin lymphoma display different clinical, morphological, immunological and molecular characteristics, allowing
them to be distinguished with reasonable precision [10, 11]. Because the two variants are also
believed to have different natural histories, they
are conventionally considered separately in epidemiological investigations whenever possible.
This is also the case in the present chapter, which
for all practical purposes will focus on classic
Hodgkin lymphoma.
Classic Hodgkin lymphomas make up around
95% of all cases and are further divided into four
subtypes referred to as nodular sclerosis (70% of
all classic Hodgkin lymphomas), mixed cellularity (20–25% of classic Hodgkin lymphomas),
lymphocyte-rich, and lymphocyte-depleted classic Hodgkin lymphoma, respectively [12, 13].
Because the distinction between the classic
Hodgkin lymphoma subtypes relies entirely on a
subjective interpretation of the histological presentation of the tumour lesion, it is subject to
1
Epidemiology of Hodgkin Lymphoma
5
both inter- and intra-observer variation [10, 11].
While classic Hodgkin lymphoma subtype may
previously have had clinical relevance, this is no
longer the case with modern imaging tools aiding
diagnosis [14].
Importantly, because of the distribution of the
Hodgkin lymphoma variants and classic Hodgkin
lymphoma subtypes, investigations reporting
associations of one kind or another for Hodgkin
lymphoma overall will still be epidemiologically
informative.
As already alluded to the discordant incidence and mortality patterns reflect that modern
therapy can cure most Hodgkin lymphoma
patients and that access to such therapy is dependent on the level of socio-economic development [16]. Of note, similar socio-economically
dependent variation in Hodgkin lymphoma mortality can also be observed in affluent countries
[17] and underscores the continued need for epidemiological investigations to promote preventive interventions.
1.3
Hodgkin Lymphoma
Occurrence
1.3.2
1.3.1
Overall Incidence
1.3.2.1 Historical Patterns
Hodgkin lymphoma occurs at all ages, but among
populations, age-specific incidence distributions
tend to vary with their ethnic and socio-economic
make-up. This variation was summarized into
four prototypical incidence patterns (numbered I
through IV) in studies in the 1950s, 1960s and
1970s [18–20]. Because they have permeated
Hodgkin lymphoma epidemiological thinking for
decades, the four patterns are briefly described
for the sake of completeness.
Pattern I was observed in developing countries
and comprised an accumulation of Hodgkin lymphoma cases—predominantly mixed cellularity—
among young boys, low incidence throughout the
second and third decades of life and increasing
incidence with age among older adults.
Pattern III was seen in affluent Western countries and demonstrated low incidence in childhood, a marked accumulation of cases—typically
nodular sclerosis—in adolescents and younger
adults (AYA), a lower incidence in middle-aged
adults and an increasing incidence with age
among older adults.
Pattern II was observed in rural areas of affluent countries and perceived as an intermediate
between patterns I and III.
Finally, a pattern IV prevailed in Asian countries and featured low incidence rates throughout
the first four decades of life followed by increasing incidence with age among older adults.
The International Agency for Research on Cancer
estimates that world-wide close to 80,000 individuals (33,431 women and 46,559 men) were
diagnosed with Hodgkin lymphoma in 2018 and
that slightly more than 26,000 individuals (10,397
women and 15,770 men) succumbed to the disease the same year [15].
This places Hodgkin lymphoma as the 27th
commonest cancer diagnosed and the 27th commonest cancer-specific cause of death [15].
Both Hodgkin lymphoma incidence and mortality display considerable geographic variation.
Overall, Hodgkin lymphoma incidence is higher
in Western world industrialized countries than in
Asian and developing countries (Fig. 1.1).
In the USA, for instance, age-standardized
(world population) incidence rates were of the
order 2.8 and 2.2 per 100,000 per year in men and
women, respectively, whereas the corresponding
figures for Indian men and women were 0.79 and
0.58 per 100,000 per year, respectively [15].
Hodgkin lymphoma mortality, conversely, is
higher in some developing countries than in
industrialized countries (Fig. 1.2). Accordingly,
age-standardized mortality rates (world) were
0.25 and 0.14 per 100,000 per year for US men
and women, respectively, and 0.53 and 0.35 per
100,000 per year for Indian men and women,
respectively [15].
Age-Specific Incidence
Patterns Vary Geographically
H. Hjalgrim and R. F. Jarrett
6
Estimated age-standardized incidence rates (World) in 2018, Hodgkin lymphoma, both sexes, all ages
ASR (World) per 100 000
≥ 2.1
1.5-2.1
0.78-1.5
0.39-0.78
< 0.39
Not applicable
No data
Fig. 1.1 Estimated age-standardized incidence rates of Hodgkin lymphoma for both sexes combined (Data from [15])
Estimated age-standardized mortality rates (World) in 2018, Hodgkin lymphoma, both sexes, all ages
ASR (World) per 100 000
≥ 0.53
0.33-0.53
0.23-0.33
0.08-0.23
Not applicable
< 0.08
No data
Fig. 1.2 Estimated age-standardized mortality rates of Hodgkin lymphoma for both sexes combined (Data from [15])
1.3.2.2 M
odern Age-Specific Incidence
Patterns
The main features of the prototypical Hodgkin
lymphoma incidence patterns, i.e. incidence
peaks in boys, AYAs and older adults, are still
recognizable.
In Chennai, India, for instance, the age-­
specific Hodgkin lymphoma incidence pattern
for the period 1993–2012 displayed a type
I-like pattern with a peak in boys less than
10 years of age, no incidence peak among adolescents and younger adults and increasing
incidence with age among older adults
(Fig. 1.3), even if a transition towards a type
II-like pattern can be observed in the more
recent of these data [15].
1
Epidemiology of Hodgkin Lymphoma
Of note, an incidence peak among boys can
also be demonstrated within European populations when more granular data are available for
analysis [21, 22].
In the USA, conversely, Hodgkin lymphoma
incidence follows a type III-like pattern with
incidence peaks among AYAs and older adults,
respectively (Figs. 1.3 and 1.4).
Even within the US age-specific incidence
patterns adhere to the correlation with socio-­
economic level. In California, a survey of cases
diagnosed 1988–1992 showed higher Hodgkin
lymphoma incidence among adolescents and
younger adults in the highest compared with
the lowest tertile of socio-economic status
(Fig. 1.4) [23].
7
Although it correlates with the level of socio-­
economic development, the geographical variation in Hodgkin lymphoma incidence likely also
reflects an association with ethnicity (Fig. 1.6).
Thus, in a Californian survey, variation in
Hodgkin lymphoma incidence rates between ethnic/racial groups was apparent even within strata
of socio-economic status [23].
1.3.3
Incidence Trends
Changes to Hodgkin lymphoma classification
systems have not been substantial in principle
allowing analyses of incidence over longer time
periods. However, considerable misclassification between Hodgkin and non-Hodgkin lym1.3.2.3 Age-Specific Incidence Patterns
phomas in the older age groups (see [10, 11, 24]
for Hodgkin Lymphoma
and references therein) resulted in inflated
Subtypes
Hodgkin lymphoma incidence rates in these age
The bimodal age distribution of Hodgkin lym- groups in earlier studies. Improvement of diagphoma incidence overall in affluent populations nostic precision may, therefore, contribute to the
is largely mirrored by the corresponding distribu- decreasing Hodgkin lymphoma incidence rates
tion of the classic Hodgkin lymphoma variants that have been reported in older age groups (e.g.
(Fig. 1.5).
[24–26]).
Both nodular sclerosis and mixed cellularity
The misclassification vis-à-vis non-Hodgkin
classic Hodgkin lymphoma display bimodal age lymphoma has been less for Hodgkin lymphoma
distributions with incidence peaks in AYA and among younger patients, rendering incidence
older adult age groups, respectively (Fig. 1.5). trend analyses more meaningful. The correlation
Mixed cellularity classic Hodgkin lymphoma between age-specific incidence patterns and level
may be the most common subtype among the of socio-economic development in the underlyyoungest children [21], but otherwise nodular ing population strongly suggests that Hodgkin
sclerosis classic Hodgkin lymphoma is the most lymphoma occurrence (and risk) is influenced by
common subtype in all age groups.
environmental factors. More specifically, it indiWhile the age-specific incidence patterns for cates that Hodgkin lymphoma incidence in
nodular sclerosis and mixed cellularity Hodgkin ­childhood and in adolescence and early adultlymphoma overall are similar for the two sexes hood is determined by correlates of socio-ecoand Hodgkin lymphoma incidence overall is nomic status or, alternatively, Westernized
higher in men than in women, the incidence of living.
classic Hodgkin lymphoma—in effect the noduTherefore, it is perhaps of little surprise that
lar sclerosis subtype—may be higher in women increasing incidence of Hodgkin lymphoma has
than in men in adolescence and early adulthood been described among adolescents and younger
(Fig. 1.3).
adults in conjunction with continued improveFor lymphocyte-depleted and lymphocyte-­ ments in living standards in both industrialized
rich classic Hodgkin lymphoma, incidence rates and developing countries (e.g. [25–29]).
generally increase with age (Fig. 1.5).
Interestingly, the rate of increase appears to have
H. Hjalgrim and R. F. Jarrett
8
Fig. 1.3 Age-specific
incidence rates for
Hodgkin lymphoma in
Chennai, India, and in
the USA in the period
1993–2012 (Data from
Bray F, Colombet M,
Mery L, Piñeros M,
Znaor A, Zanetti R and
Ferlay J, editors (2017)
Cancer Incidence in Five
Continents, Vol. XI
(electronic version)
Lyon, IARC. http://ci5.
iarc.fr last accessed on
[25 March 2019])
Hodgkin Iymphoma (1993-2012)
6.5
6.0
5.5
5.0
4.5
rate per 100,000
4.0
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
0-
10-
20-
30-
4050age
60-
70-
80-
International Agency for Research on Cancer (IARC) - 25.3.2019
USA, SEER (9 registries) : Male
USA, SEER (9 registries) : Female
India, Chennai : Male
India, Chennai : Female
been more pronounced among AYA women than
among AYA men.
Because diagnostic misclassification also
extends to subtypes of classic Hodgkin lymphoma [10, 11] and because the increasing use of
non-excisional biopsies for lymphoma diagnosis
limits the tumour material available for diagnostic purposes [30], changes in classic Hodgkin
lymphoma subtype-specific trends are also difficult to interpret.
Epidemiology of Hodgkin Lymphoma
Fig. 1.4 Age-specific
incidence rates of
Hodgkin lymphoma by
tertile of neighbourhood
socio-economic status,
California, 1988–1992
(Figure reproduced from
[23])
9
7
6
5
Rate/100,000
1
4
3
2
1
0
0-4
10-14
20-24
30-34
40-44
50-54
60-64
70-74
80-84
Age at diagnosis
Medium
High
Low
3.5
Incidence per 100,000 per year
3
2.5
2
1.5
1
0.5
0
00–09 yr
10–19 yr
20–29 yr
30–39 yr
40–49 yr
50–59 yr
60–69 yr
70–79 yr
80+ yr
Age group
Nodular sclerosis
Mixed cellularity
Lymphocyte depleted
Lymphocyte rich
Not otherwise specified
Fig. 1.5 Age-specific incidence rates of classic Hodgkin lymphoma in both sexes combined by histological subtype in
the USA (2000–2011) (Data from US SEER 18. This Figure was reproduced from [6])
H. Hjalgrim and R. F. Jarrett
10
7
Whites
Blacks
Hispanics
Asian/Pacific Islanders
Incidence per 100,000 per year
6
5
4
3
2
1
0
00–09 yr
10–19 yr
20–29 yr
30–39 yr
40–49 yr
50–59 yr
60–69 yr
70–79 yr
80+ yr
Age group
Fig. 1.6 Age-specific incidence rates of classic Hodgkin lymphoma overall in both sexes combined by race/ethnicity
in the USA (2008–2012) (Data from US SEER 18. Figure reproduced from [6])
1.3.4
Classifications for
Epidemiological Studies:
Multi-disease Models
Efforts to unravel causes of Hodgkin lymphoma
have been complicated by the strong suspicion
that several epidemiologically and etiologically
distinct Hodgkin lymphoma variants exist and
because efforts to define these have proven
exceedingly difficult.
1.3.5
lassifications by Age at
C
Diagnosis, Histology and
Tumour Epstein-Barr Virus
Status
Intrigued by the bimodal age distribution and by
corresponding epidemiological and clinical variation between cases within the age-specific incidence
peaks, MacMahon in 1966 [19] proposed that three
etiologically heterogeneous Hodgkin lymphoma
types existed and that age at diagnosis, specifically
0–14 years, 15–34 years and 50+ years, could be
used as a proxy to distinguish between them.
MacMahon, moreover, suggested that
Hodgkin lymphoma in young adults had an infectious aetiology [19].
As information on Hodgkin lymphoma subtypes, classified using modern criteria, became
available for larger patient series, the composition of cases within the incidence peaks (Fig. 1.6)
led to the proposal that nodular sclerosis and
mixed cellularity Hodgkin lymphoma each
­captured one of the supposedly etiologically distinct variants.
In 1985, Poppema and colleagues were the
first to report the presence of Epstein-Barr
virus genome products in the malignant
Hodgkin/Reed-Sternberg cells in a patient with
Hodgkin lymphoma [31]. Nearly 35 years later,
we now know that some 30% of Hodgkin lymphomas among adults in affluent Western populations and even more in African and Asian
countries are positive for Epstein-Barr virus
[32]. We also know that epidemiologically
Epstein-Barr virus-­
positive and Epstein-Barr
virus-negative Hodgkin lymphomas differ
(reviewed in [33], and further discussed in
Chap. 2). Consequently, tumour Epstein-Barr
1
Epidemiology of Hodgkin Lymphoma
11
virus status offers itself as a third way to group
classic Hodgkin lymphoma into epidemiologically distinct entities.
1.3.6
miologically specific entities has empirical merit,
as will be discussed in the sections below.
However, they are sufficiently incongruent to
reflect the same phenomenon or subtype.
Age at diagnosis displays some overlap with
both histology (Fig. 1.5) and tumour Epstein-­
Barr virus status (Fig. 1.7). However, there is
less overlap between tumour Epstein-Barr virus
status and histological subtype. While most
mixed cellularity classic Hodgkin lymphomas
are Epstein-Barr virus-positive and most nodular
verlap Between
O
Epidemiological Classifications
of Hodgkin Lymphoma
Each of the three proposed means to stratify
Hodgkin lymphoma into etiologically and epidea
6
5
Rate/100,000
4
3
2
1
0
20
30
40
50
60
70
Age (years)
b
2.5
2.0
Rate/100,000
Fig. 1.7 Age-specific
incidence rates for
classic Hodgkin
lymphoma overall
(dotted line), Epstein-­
Barr virus-negative
(squares) and virus-­
positive (triangles)
classic Hodgkin
lymphomas. Panel b:
Age-specific Epstein-­
Barr virus-positive
classic Hodgkin
lymphoma among men
(squares) and women
(triangles) (Figure
reproduced from [10])
1.5
1.0
0.5
0
20
30
40
50
Age (years)
60
70
H. Hjalgrim and R. F. Jarrett
12
sclerosis classic Hodgkin lymphomas are
Epstein-Barr virus-negative, nodular sclerosis
classic Hodgkin lymphomas still make up more
than half of all Epstein-Barr virus-positive
cases [33].
Because Epstein-Barr virus remains the most
plausible causal candidate for Hodgkin lymphoma, its incongruence with the tumour histology classification is insufficient grounds for its
dismissal. Therefore, age at diagnosis, tumour
histology and tumour Epstein-Barr virus status
likely represent different elements of Hodgkin
lymphoma natural history, emphasizing its
complexity.
1.4
Familial Accumulation
of Hodgkin Lymphoma:
Genetic Predisposition
It has been known for more than half a century
that Hodgkin lymphomas cluster within families,
the earliest investigation indicating an approximately threefold increased risk among Hodgkin
lymphoma patients’ first-degree relatives [34].
Subsequent studies with access to larger and
even register-based data have largely confirmed
this association. The hitherto largest investigation, a Nordic register-based investigation of
57,475 first-degree relatives of 13,922 classic
Hodgkin lymphoma patients, yielded an overall
standardized incidence ratio of 3.3 (95% confidence interval 2.8–3.9) for familial recurrence,
corresponding to a 0.6% lifetime risk of classic
Hodgkin lymphoma among patient relatives [35].
Owing to its magnitude, the Nordic study
allowed for stratification of analyses by type of
family relation and found a standardized incidence
ratio of 2.1 (95% confidence interval 1.6–2.6) for
the combined group of patient parents and offspring, but 6.0 (95% confidence interval 4.8–7.4)
among patient siblings [35]. Even more extremely
increased risks were observed for same-sex twins
(standardized incidence ratio 57 (95% confidence
interval 21–125)), consistent with the results of a
previous American twin study [36].
Hodgkin lymphoma also clusters with other
haematological malignancies, notably chronic
lymphocytic leukaemia [37] and diffuse large
B-cell lymphoma [38], with relative risks for
these malignancies being slightly less increased—
twofold—than for classic Hodgkin lymphoma.
1.4.1
enetic Studies: Genome-­
G
Wide Association Studies
The accumulation of Hodgkin lymphomas among
relatives must reflect shared environmental and
constitutional risk factors. Both candidate gene
investigations and genome-wide association
studies confirm the suspicion that genetic predisposition is important to Hodgkin lymphoma risk.
The histological presentation of Hodgkin lymphoma is dominated by admixture of accessory
and inflammatory cells indicating that immune
function is important to the disease. Early
research, therefore, focused on the association
between tissue type, i.e. HLA, and Hodgkin lymphoma risk. Indeed, Hodgkin lymphoma was
among the first diseases linked with markers of
specific HLA types [39, 40].
Genotyping of more than 5000 patients with
Hodgkin lymphoma has identified a total of 18
genetic loci associated with Hodgkin lymphoma
risk. Based on these loci, Sud and colleagues
point to three key biological processes underlying
Hodgkin lymphoma susceptibility. These are (1)
the germinal centre reaction (2p16.1, REL; 3q28,
BCL6 and mir-28; 6p21, HLA; 6q23.3, MYB;
8q24.21, MYC; 11q23.1, POU2AF1; 16p11.2,
MAPK3; 19p13.3, TCF3; 20q13.12, CD40), (2)
T-cell differentiation and function (3p24.1,
EOMES; 5q31,1, IL13; 6q22.33, PTPRK and
THEMIS; 6q23.3, MYB; 6q23.3, AHI1; 10p14,
GATA3; 16p13.1, SOCS1 and CLEC16A;
16p11.2, MAPK3 and CORO1A) and (3) constitutive NF-kB activation (2p16.1, REL; 3p24.1,
AZI2; 6q23.3, TNFAIP3; 20q13.12, CD40) [41].
1.4.1.1 H
odgkin Lymphoma Subtype-­
Specific Associations in Genetic
Analyses
In addition to shedding light on possible mechanisms underpinning Hodgkin lymphoma pathogenesis in general (discussed in the following
1
Epidemiology of Hodgkin Lymphoma
chapter (The Role of Viruses in the Genesis of
Hodgkin Lymphoma)), the genetic investigations also add further credence to the notion of
etiological heterogeneity between classic
Hodgkin lymphoma subtypes.
Accordingly, as discussed in the following
chapter, evidence is mounting that risk of EpsteinBarr virus-positive Hodgkin lymphoma is
strongly associated with alleles in the HLA class
I region (e.g. rs2734986, HLA-A; rs6904029,
HCG9), whereas risk of Epstein-Barr virus-negative disease is more strongly associated with
alleles within the HLA class II region (e.g.
rs6903608, HLA-DRA) [42].
The variation in genetic associations underscores the importance of detailed phenotyping in
genome-wide association as well as other types
of studies of Hodgkin lymphoma. Thus, in the
absence of information on either tumour histology or tumour Epstein-Barr virus status, neither
the presence nor absence of associations can be
fully interpreted.
This is illustrated by an extended analysis of
the HLA region in the first GWAS to stratify
cases by EBV status [42, 43]. This study
revealed a single nucleotide polymorphism near
the HLA-­DPB1 gene (rs6457715) which was
associated with Epstein-Barr virus-positive
(odds ratio 2.33 (95% confidence interval 1.83–
2.97; P 10–12)), but not with Epstein-Barr
virus-negative Hodgkin lymphoma risk (odds
ratio 1.06 (95% confidence interval 0.92–1.21);
Phom = 10−8), a difference that was present even
within strata defined by classic Hodgkin lymphoma histology [43].
With that reservation, the association with
class I alleles for Epstein-Barr virus-positive
Hodgkin lymphomas or with mixed cellularity
Hodgkin lymphoma suggests that cytotoxic
T-cells’ control of virally infected lymphocytes
whether before or after malignant transformation
is significant to the risk of the lymphoma.
Epstein-Barr virus-negative Hodgkin lymphoma’s association with HLA class II alleles may
reflect a similar role of immune control of an
infectious agent in the pathogenesis of this
Hodgkin lymphoma subtype. Alternatively, it
may also be indicative of the involvement of
13
CD4-positive T follicular helper cells in Hodgkin
lymphoma pathogenesis [44].
For further discussion, please be referred to
Chap. 2.
1.5
Risk Factors
1.5.1
Prevailing Hypotheses
in Hodgkin Lymphoma
Epidemiology
As already mentioned, for more than half a century, it has been assumed that Hodgkin lymphomas in children, AYA and older adults differ
epidemiologically—possibly aetiologically—
from one another [19]. Therefore, epidemiological investigations have conventionally
considered risk factors for Hodgkin lymphoma
for the three age groups separately, when practically possible.
Owing to the bimodal age distribution of
Hodgkin lymphoma in affluent populations, the
epidemiology of AYA Hodgkin lymphoma has
been the most studied.
1.5.1.1 Childhood SocioEconomic Environment
The correlation between age-specific Hodgkin
lymphoma incidence patterns and level of
socio-­economic development in the underlying
population led to the formulation of the socalled late infection model for Hodgkin lymphoma [45, 46].
This model suggests that Hodgkin lymphoma
among children, adolescents and younger adults
is caused by an infectious agent and that lymphoma risk increases with increasing age at primary infection [45, 46].
Early studies supported this understanding of
AYA Hodgkin lymphoma indirectly by reporting
that correlates of childhood socio-economic
affluence, which are thought to be associated
with low childhood infectious disease pressure,
such as length of maternal education, home ownership and being a member of a small sibship,
were associated with increased risk of AYA
Hodgkin lymphoma [47]. Moreover, within sib-
H. Hjalgrim and R. F. Jarrett
14
ships, Hodgkin lymphoma risk correlated
inversely with number of older siblings, i.e. birth
order [47].
Mack and colleagues recently (2015) reported
the results of a register-based case-control study
nested in a cohort of US army conscripts in the
period 1950–1968. The study included 656 men
diagnosed with Hodgkin lymphoma at ages
17–32 years and individually matched controls.
In univariate analyses, they found increased
Hodgkin lymphoma risk with small sibship size
(odds ratio 1.4 (95% confidence interval 1.1–1.9)
for 2–3 vs. >3 children), birth order (odds ratio
1.9 (95% confidence interval 1.4–2.6) for first vs.
middle born) and short age gap to nearest sibling
(odds ratio 2.1 (95% confidence interval 1.5–3.1)
for more vs. less than 5 years) [48].
Information on histological subtype was available for a subset of the cases in the study of US
conscripts, but analyses did not point to different
associations for nodular sclerosis and mixed cellularity Hodgkin lymphoma. No information on
tumour Epstein-Barr virus status was available
for analyses [48].
In later investigations—that is, studies of
patients diagnosed in more recent years—the
association between traditional measures of
childhood socio-economic affluence and AYA
Hodgkin lymphoma risk in Western countries has
been less compelling or even absent [49–54].
This change in the epidemiology of Hodgkin
lymphoma in AYA may suggest that family or
rather sibship structure no longer reflects the
early life exposures associated with the disease
[51]. Of note in this regard, in two case-control
investigations, one American and one
Scandinavian, kindergarten attendance was associated with reduced Hodgkin lymphoma risk in
AYAs (odds ratio 0.64 (95% confidence interval
0.45–0.92) and odds ratio = 0.78 (95% confidence interval 0.56–1.09)) [49, 52].
Information on tumour Epstein-Barr virus status was available in both investigations, and while
no differences in association with nursery school
attendance were observed in the American investigation, it tended to be stronger for virus-­negative
Hodgkin lymphoma in the Scandinavian study
[49, 52]. This contrast highlights that varying
associations may simply result from the subtype
composition of the studied Hodgkin lymphomas.
The evidence supporting the idea that childhood socio-economic environment influences
Hodgkin lymphoma risk in childhood is also
insubstantial. One Danish register cohort study
found that risk of Hodgkin lymphoma before age
15 years increased with sibship size and birth
order [55]. This association was reproduced in
neither Swedish [56] nor American data [57].
However, a large North American case-control
investigation reported findings similar to the
Danish study: increasing sibship size was positively and increasing maternal education and
household income inversely associated with
Hodgkin lymphoma risk before age 15 years [58].
1.5.2
Anthropometry
Hodgkin lymphoma risk up to the age of early
adulthood in some investigations (though not all)
associates with increasing birth weight (e.g. [53,
54]). In a Californian register-based investigation,
Hodgkin lymphoma risk in the age group
0–19 years was found to increase by 16% (95%
confidence interval 1.03–1.30) per kilogram
increase in birth weight [54]. While in this investigation the association appeared specific to nodular
sclerosis classic Hodgkin lymphoma [54], it
applied to both nodular sclerosis and mixed cellularity classic Hodgkin lymphoma in the other [53].
Like reports of association between stature
late in childhood/in early adolescence and subsequent risk of Hodgkin lymphoma [59, 60], the
association with birth weight may at least in part
reflect Hodgkin lymphomas association with
childhood socio-economic affluence.
A number of prospective investigations have
pointed to an association between obesity and
Hodgkin lymphoma risk [61–63]. For example,
in the UK Million Women Study, body mass
index correlated with Hodgkin lymphoma risk
(hazard ratio 1.64 (95% confidence interval 1.21–
2.21) per 10 kg per square meter increase)) [63].
If true, the association between obesity and
Hodgkin lymphoma risk could reflect obesity-­
related inflammation references.
1
Epidemiology of Hodgkin Lymphoma
1.5.3
Medical History
15
extension of the Scandinavian cohort study mentioned above, according to which infectious
mononucleosis was associated with an increased
1.5.3.1 Infections
The late infection model fostered much interest risk of Epstein-Barr virus-positive classic
in the search for infectious agents that might Hodgkin lymphomas (standardized incidence
cause Hodgkin lymphoma. Among suspected ratio 4.0 (95% confidence interval 3.4–4.5)) and
organisms, the human herpesvirus Epstein-Barr not Epstein-Barr virus-negative classic Hodgkin
virus, first isolated from Burkitt lymphoma tissue lymphoma (standardized incidence ratio 1.5
[64] and soon after established as the cause of (95% confidence interval 0.9–2.5)) [68].
While similar subtype-specific observations
infectious mononucleosis [65], has long been the
were also made in British and Scandinavian case-­
centre of attention.
control investigations [49, 69], other studies have
reported increased risks for both Epstein-Barr
Epstein-Barr Virus Infection: Infectious
virus-positive and Epstein-Barr virus-negative
Mononucleosis
Epidemiological, serological and molecular-­ Hodgkin lymphomas [70] or no associations at
biological (i.e. the presence of Epstein-Barr virus all [50, 52].
genome products in the malignant cells) evidence
link Epstein-Barr virus infection to Hodgkin Epstein-Barr Virus Infection: Serological
lymphoma development. Here, only the epide- Evidence
miological and serological evidence will be pre- Support for the association between Epstein-Barr
sented; for the presence and role of Epstein-Barr virus infection and Hodgkin lymphoma risk also
comes from serological investigations [71, 72].
virus in the malignant cells, please see Chap. 2.
In 1989 Nancy Mueller and colleagues in a
Infectious mononucleosis is rarely seen
prospective
nested case-control study found that
among children but is a common presentation of
primary infection with the Epstein-Barr virus aberrant patterns of anti-Epstein-Barr virus antiwhen it is delayed until adolescence [66]. bodies were associated with overall Hodgkin
Numerous investigations have assessed the asso- lymphoma risk [73].
Three decades later this study was replicated
ciation between infectious mononucleosis and
Hodgkin lymphoma, and most have reported only this time with information on tumour
increased Hodgkin lymphoma risk in the wake of Epstein-Barr virus status. Comparing pre-­
diagnostic antibody patterns in 40 and 88
infectious mononucleosis (reviewed in [33]).
The largest of these was a Scandinavian patients with Epstein-Barr virus-positive and
Barr virus-negative classic Hodgkin
register-­based cohort study of more than 40,000 Epstein-­
patients with infectious mononucleosis followed lymphomas with those in matched controls,
for the occurrence of Hodgkin lymphoma. Levin and colleagues showed that an inverted
EBNA2 antibody level ratio
Compared with the general population, the infec- anti-EBNA1/anti-­
tious mononucleosis patients were at a 2.55 (95% (≤1) consistent with impaired control of
confidence interval 1.87–3.40)-fold increased Epstein-Barr virus infection was associated with
Hodgkin lymphoma risk. The risk increase was a 4.7 (95% confidence interval 1.6–13.8)-fold
particularly high shortly after the Epstein-Barr increased risk of Epstein-Barr virus-positive
virus infection but remained increased for up to Hodgkin lymphoma, whereas no association
negative
20 years of follow-up [67]. Because infectious was observed for Epstein-Barr virus-­
mononucleosis typically occurs in adolescence, Hodgkin lymphoma [74].
the increased Hodgkin lymphoma risk tended to
Epstein-Barr Virus Infection: Variation
present in younger adults.
In a few investigations, information on in Tumour Prevalence
Hodgkin lymphoma Epstein-Barr virus status has Epstein-Barr virus can, as already mentioned, be
been available for analyses. One such was an demonstrated in the malignant cells in a proportion
16
of classic Hodgkin lymphomas [32]. Importantly,
however, its presence is non-­randomly distributed
between cases and tends to reflect ethnic, sociodemographic, age, sex and disease-specific circumstances, adding further support to the suspicion of a
causal relation.
This variation was most eloquently demonstrated by Glaser and colleagues in a pooled analysis of 1546 patients [75]. The analysis showed
that irrespective of age at diagnosis, Epstein-Barr
virus could more often be demonstrated in mixed
cellularity than in nodular sclerosis Hodgkin
lymphoma, in children from deprived rather than
affluent settings and in male than in female
patients, except among older adults. At the same
time, compared with adolescents and younger
adults, Hodgkin lymphomas in children and older
adults were more often Epstein-Barr virus-­
positive [75].
The seminal paper by Glaser and colleagues
once again underscores the importance of information on histological subtype and tumour
Epstein-Barr virus status in epidemiological
investigations.
H. Hjalgrim and R. F. Jarrett
phoma has so far been in vain (see also Chap. 2).
Indeed, direct support for the decreased infectious disease load in early childhood among adolescent and younger adult Hodgkin lymphoma
patients implied by the late infection hypothesis
is scarce.
Even if in interview-based case-control studies
self-reported history of infections such as measles,
mumps and rubella have been associated with
reduced risk of Hodgkin lymphoma in adolescence
and early adulthood [50, 69, 77, 78], the validity of
such recalled childhood health history may be
questioned. Still, in Mack and colleagues’ prospective study of army conscripts, history of mumps
ascertained at the start of follow-­up also was associated with reduced Hodgkin lymphoma risk [48].
Cozen and colleagues retrospectively assessed
childhood exposures likely to produce oral exposure to microbes among 188 sets of twins discordant for Hodgkin lymphoma diagnosed at ages
13–50 years. Most interestingly, their study
showed that Hodgkin lymphoma risk was lower
for the twins whose behaviour mostly likely led
to oral microbial exposure [79].
A Four-Disease Model for Hodgkin
1.5.3.2 Primary and Secondary
Immune Deficiencies
Lymphoma
The age-dependent variation in prevalence of Similar to non-Hodgkin lymphomas, Hodgkin
Epstein-Barr virus-positive Hodgkin lymphomas lymphomas also occur excessively among
along with its association with infectious mono- patients suffering from (certain) primary and secnucleosis has given rise to the four-disease model, ondary immune deficiencies (reviewed in [80]).
according to which Epstein-Barr virus-positive
Risk of Hodgkin lymphoma is between 4- and
and Epstein-Barr virus-negative Hodgkin lym- 16-fold increase in cohort studies of HIV-infected
phomas are etiologically separate entities [76]. people and between 2- and 7-fold increase in
The model is further discussed in Chap. 2, but in cohort studies of solid organ transplant recipients
summary suggests that Epstein-Barr virus-­ (reviewed in [80]).
positive Hodgkin lymphoma develops in conMost Hodgkin lymphoma occurring in the setjunction to primary infection with the virus ting of immune deficiency is Epstein-Barr virus-­
(children and adolescents) or because of subse- positive [80]. Correspondingly, in one cohort
quent loss of control with the viral infection in its study of people with AIDS-related immune defichronic phase owing to immune impairment for a ciency, risk was more increased for mixed celluvariety of reasons, while no causes for Epstein-­ larity (rate ratio 18.3 (95% confidence interval
Barr virus-negative Hodgkin lymphoma are 15.9–20.9)) and lymphocyte-depleted (rate ratio
suggested.
35.3 (95% confidence interval 24.7–48.8))
Hodgkin lymphoma subtypes than for nodular
Other Childhood Infections
sclerosis Hodgkin lymphoma [81].
The search for other specific childhood infections
There is (some) evidence that Hodgkin lymcausally associated with classic Hodgkin lym- phoma risk correlates inversely with degree of
1
Epidemiology of Hodgkin Lymphoma
immune suppression as measured by CD4-­
lymphocyte count among HIV-infected people,
but the correlation is less strong than for non-­
Hodgkin lymphoma and, in contrast to the latter,
risk does not correlate with measures of HIV
load [82, 83].
These differences between Hodgkin and non-­
Hodgkin lymphoma may explain why the overall
incidence of Hodgkin lymphoma has not decreased
to the same extent as non-Hodgkin lymphoma following the introduction of highly active antiviral
therapy (see [82] and references therein).
1.5.3.3 Autoimmune and Allergic
Disorders
Autoimmune and Allergic/Atopic Diseases
Several studies have reported an increased risk of
Hodgkin lymphoma among patients with autoimmune diseases (see reviews [80, 84]). A large
Swedish investigation reported a twofold increased
risk of Hodgkin lymphoma risk among 878,000
patients registered with any of 33 autoimmune
conditions in the Swedish inpatient register [85].
Temporal variation in relative risk of Hodgkin
lymphoma suggested that the increased risk was
partially explained by reversed causality; specifically, that incipient (undiagnosed) Hodgkin lymphoma led to autoimmune disease diagnosis.
Thus, the standardized incidence ratio (SIR) for
Hodgkin lymphoma decreased with time since
autoimmune disease diagnosis from 5.2 (95%
confidence interval 4.2–6.3) in the first year of
follow-up to 2.0 (95% confidence interval 1.7–
2.4) in the period 1–4 years after autoimmune disease diagnosis and to 1.5 (1.2–1.7) at 5 or more
years after autoimmune disease diagnosis [85].
Accordingly, when the first 5 years after autoimmune diagnosis was disregarded, statistically
significantly increased risk of Hodgkin lymphoma was observed for patients with rheumatoid arthritis (SIR = 2.0 (95% confidence interval
2.1–3.7)), autoimmune haemolytic anaemia
(SIR = 16.6 (95% confidence interval 3.1–49.2)),
Behcet disease (SIR = 4.0 (95% confidence interval 1.3–9.3)) and systemic lupus erythematosus
(SIR = 4.1 (95% confidence interval 1.5–9.0)).
Elevated risk estimates, albeit not statistically
17
significant, were still observed for various other
autoimmune diseases [85].
The mechanisms underlying the association
between Hodgkin lymphoma and autoimmune
diseases have remained elusive. Interestingly, in
another register-based study from Sweden,
reduced risk of Hodgkin lymphoma was observed
in families of patients with acute glomerular
nephritis, ankylosing spondylitis and Graves disease and increased in family members of patients
with pemphigus. In first-degree relatives of
Hodgkin lymphoma patients, several autoimmune diseases occurred in excess (Behcet disease, dermatitis herpetiformis, multiple sclerosis,
primary biliary cirrhosis and rheumatoid arthritis) or in deficit (celiac disease and psoriasis)
pointing to some form of shared risk between the
two disease groups [86], which to some extent
could be genetic in nature [87].
Multiple sclerosis stands out from other autoimmune diseases in this regard. Thus, studies
have shown that Hodgkin lymphoma in young
adults and multiple sclerosis clusters mutually
within individuals [88] and within families [86,
89]. Moreover, the two conditions have also been
found to share genetic risk profiles to the extent
that polygenic risk scores for either of the two
diseases are associated with risk of the other [87].
Interestingly, these two disparate conditions both
share the association with infectious mononucleosis, raising the possibility that the three conditions are somehow immunologically related.
Few studies have examined the association
between allergic/atopic diseases and Hodgkin
lymphoma risk ([90], review in [91]).
Methodological issues aside, the results of the
analyses are too heterogeneous to preclude conclusions other than that the current evidence does
not support any association between the two. In
agreement with this, Levin and colleagues found
no evidence of an association between pre-­
diagnostic titres of IgE and Hodgkin lymphoma
risk in a prospective serological investigation [92].
1.5.3.4 Medications
Regular use of aspirin has been suggested to be
associated with reduced Hodgkin lymphoma
risk in one American [93] and in two partly
H. Hjalgrim and R. F. Jarrett
18
overlapping Danish studies [94]. Combining the
two sets of results in a meta-analysis, long-term
use of aspirin was associated with an odds ratio
of 0.62 (95% confidence interval 0.46–0.82) for
Hodgkin lymphoma [94]. Aspirin’s interference
with inhibition of NF-κB which is constitutively
active in the malignant Hodgkin/Reed-Sternberg
cells and its binding to cyclooxygenase (COX)-1
and cyclooxygenase-2, which are overexpressed
in Hodgkin lymphoma, both lend biological
plausibility to the observed association [93, 94].
Of note, the same investigations also reported
increased Hodgkin lymphoma risk among users
of other nonsteroidal anti-inflammatory drugs
such as selective Cox-2 inhibitors or acetaminophen. However, temporal variation in the risk
increase suggested that reverse causality—confounding by indication—most likely accounted
for the observed association [93, 94].
1.5.4
Environmental Exposures
1.5.4.1 Ultraviolet Light
Recent decades have witnessed considerable epidemiological interest in the possible association
between vitamin D and cancer because of the
vitamin’s potential anticarcinogenic effects [95].
Because ultraviolet light radiation is critical to
vitamin D production, studies have typically
focused on this exposure, as is also the case for
Hodgkin lymphoma investigations.
An association between ultraviolet radiation
exposure and Hodgkin lymphoma risk is supported by two types of evidence. Firstly, according to ecological studies, Hodgkin lymphoma
incidence rates and ambient levels of ambient
ultraviolet radiation correlate inversely [96, 97].
Secondly, according to a pooled analysis of
data from four case-control studies including a
total of 1320 patients with Hodgkin lymphoma
and 6381 controls, history of sunburn (odds
ratio = 0.77 (95% confidence interval 0.63–0.95))
and sunlamp use (odds ratio = 0.81 (95% confidence interval 0.69–0.96)) and cumulative lifetime exposure to ultraviolet radiation were each
associated with statistically significantly
decreased risk of Hodgkin lymphoma [98]. The
observed associations tended to be stronger for
Epstein-Barr virus-positive than for Epstein-Barr
virus-­negative Hodgkin lymphomas [98].
Both the ecological and the analytical epidemiological data are compatible with ultraviolet
radiation exposure in some way or other preventing Hodgkin lymphoma pathogenesis.
1.5.4.2 Tobacco
Considering tobacco’s many effects on the human
immune system, it is conceivable that it is also
associated with Hodgkin lymphoma risk [99].
Support for this notion comes from two meta-­
analyses and a pooled analysis of several large
datasets.
The meta-analyses both conclude that current
smoking carries a statistically significant 30–40%
increased risk of Hodgkin lymphoma [100, 101],
whereas the pooled analyses suggest a statistically nonsignificant 16% increased risk [102].
Both meta-analyses also find evidence of doseresponse associations between Hodgkin lymphoma
risk and current cigarette smoking measured as
number of cigarettes smoked, years smoked and
pack-years [100, 101]. The pooled analysis, in contrast, found no evidence of a dose-­response pattern
in the association between Hodgkin lymphoma risk
and cigarette smoking [102].
In stratified analyses, the association with current smoking was stronger for mixed cellularity
than for nodular sclerosis Hodgkin lymphoma
[100] and correspondingly also stronger for
Epstein-Barr virus-positive than for Epstein-Barr
virus-negative Hodgkin lymphoma [101, 102].
1.5.4.3 Alcohol
A reduced Hodgkin lymphoma risk with alcohol
intake has been suggested by most investigations
of the topic, even if observed associations do not
always reach statistical significance and rarely
display dose-response patterns [103–112].
In a recent meta-analysis of available cohort
studies, ever drinking alcohol was associated
with a statistically nonsignificant reduced
Hodgkin lymphoma risk (relative risk = 0.74
(95% confidence interval 0.52–1.05)) [113].
Although results vary between studies, the
reported inverse association has been reported for
1
Epidemiology of Hodgkin Lymphoma
19
all Hodgkin lymphoma subgroups, i.e. for both perspective of the understanding of what drives
younger and older adult patients, for Epstein-­ Hodgkin lymphoma development, this represents
Barr virus-positive and Epstein-Barr virus-­ a field of research that has only been little explored
negative Hodgkin lymphoma, as well as for in Epstein-Barr virus status-­specific contexts.
nodular sclerosis and mixed cellularity Hodgkin
The favourable prognosis of Hodgkin lymlymphoma alike.
phoma achievable with modern therapy is
One caveat to the interpretation of the observed unlikely to foster clinical interest into the clinical
reduced Hodgkin lymphoma risk is that the lym- significance of tumour Epstein-Barr virus status
phoma may be accompanied by alcohol intoler- or other (potential) markers of baseline treatment
ance [114]. Consequently, reduced alcohol intake needs. Accordingly, though it could further
could result from early Hodgkin lymphoma man- access to large dataset amenable for research, the
ifestations. An attempt to mitigate this problem motivation to determine epidemiologically relewas introduced in the UK Million Study in which vant markers in clinical trials is modest.
an increased risk of Hodgkin lymphoma among
never drinkers (hazard rate ratio = 1.70 (95%
confidence interval 1.27–2.26)) compared with References
occasional drinkers (0.5–3 drinks weekly) was
1. Engert A, Haverkamp H, Kobe C et al (2012)
unaffected when the first 3 years of follow-up
Reduced-intensity chemotherapy and PET-guided
was disregarded [112]. Still, this study also
radiotherapy in patients with advanced stage
showed no dose-response pattern between alcoHodgkin’s lymphoma (HD15 trial): a randomised,
open-label, phase 3 non-inferiority trial. Lancet
hol consumption and Hodgkin lymphoma risk.
1.6
Conclusion
Studies have demonstrated that Hodgkin lymphomas occurring at different ages have different
epidemiologic profiles. This variation is commonly interpreted as evidence that Hodgkin lymphoma comprises two or more aetiologically
heterogeneous conditions.
Age, histological presentation and tumour
Epstein-Barr virus status have been suggested to
identify unique classic Hodgkin lymphoma entities. Evidence is strong that Epstein-Barr virus-­
positive Hodgkin lymphomas are aetiologically
different from their Epstein-Barr virus-negative
counterparts. There is also good evidence that the
risk of Epstein-Barr virus-positive Hodgkin lymphoma at different ages to a large extent is influenced by circumstances influencing age at primary
infection and immunological response to or control of the infection, whether in its acute or chronic
phases. Meanwhile, the causes of Epstein-Barr
virus-negative Hodgkin lymphoma have remained
elusive and call for continued research.
Both Epstein-Barr virus-positive and Epstein-­
Barr virus-negative Hodgkin lymphoma can have
different histopathological presentations. From the
379(9828):1791–1799
2. Engert A, Plütschow A, Eich HT et al (2010)
Reduced treatment intensity in patients with
early-stage Hodgkin’s lymphoma. N Engl J Med
363(7):640–652
3. Hodgkin T (1832) On some morbid appearances of
the absorbent glands and spleen. Med Chirurg Trans
17:68–114
4. Ng AK, van Leeuwen FE (2016) Hodgkin lymphoma: late effects of treatment and guidelines for
surveillance. Semin Hematol 53(3):209–215
5. Cohen JI (2018) Vaccine development for Epstein-­
Barr virus. Adv Exp Med Biol 1045:477–493
6. Hjalgrim H, Chang ET, Glaser SL (2018) Hodgkin
lymphoma. Schottenfeld and Fraumeni Cancer
Epidemiology and Prevention. Oxford University
Press:745–766
7. Stein H (2001) Hodgkin lymphomas: introduction. WHO Classif Tumours Tumours Haematop
Lymphoid Tissues 8:239
8. Stein H, Delsol G, Pileri S et al (2001) Nodular
lymphocyte predominant Hodgkin lymphoma.
WHO Classif Tumours Tumours Haematopoietic
Lymphoid Tissues:240–243
9. Stein H, Delsol G, Pileri S et al (2001, 2001)
Classical Hodgkin lymphoma. WHO Classif
Tumours Tumours Haematopoietic Lymphoid
Tissues:244–253
10. Jarrett RF, Krajewski AS, Angus B et al (2003) The
Scotland and Newcastle epidemiological study of
Hodgkin’s disease: impact of histopathological
review and EBV status on incidence estimates. J
Clin Pathol 56(11):811–816
11. Glaser SL, Dorfman RF, Clarke CA (2001) Expert
review of the diagnosis and histologic classification
H. Hjalgrim and R. F. Jarrett
20
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
of Hodgkin disease in a population-based cancer
registry: interobserver reliability and impact on incidence and survival rates. Cancer 92(2):218–224
Stein H, von Wasielewski R, Poppema S, MacLennan
K, Guenova M (2008) Nodular sclerosis classical Hodgkin lymphoma. WHO Classif Tumours
Tumours Haematopoietic Lymphoid Tissues
2008:330
Weiss LM, von Wasielewski R, Delsol G, Poppema S,
Stein H (2008) Mixed cellularity classical Hodgkin
lymphoma. WHO Classif Tumours Tumours
Haematopoietic Lymphoid Tissues 2008:331
Ansell SM (2015) Hodgkin lymphoma: diagnosis
and treatment. Mayo Clin Proc 90(11):1574–1583
Ferlay J, Ervik M, Lam F, Colombet M, Mery L,
Piñeros M, Znaor A, Soerjomataram I, Bray F
(2018). Global Cancer Observatory: Cancer Today.
Lyon, France: International Agency for Research
on Cancer. Available from: https://gco.iarc.fr/today,
accessed [25 March 2019].
Chatenoud L, Bertuccio P, Bosetti C et al (2013)
Hodgkin’s lymphoma mortality in the Americas,
1997–2008: achievements and persistent inadequacies. Int J Cancer 133(3):687–694
Keegan THM, DeRouen MC, Parsons HM et al
(2016) Impact of treatment and insurance on socioeconomic disparities in survival after adolescent
and young adult Hodgkin lymphoma: a population-­
based study. Cancer Epidemiol Biomark Prev
25(2):264–273
Macmahon B (1958) Epidemiological evidence of the nature of Hodgkin’s disease. Cancer
10(5):1045–1054
MacMahon B (1966) Epidemiology of Hodgkin’s
disease. Cancer Res 26(6):1189–1201
Correa P, O’Conor GT (1971) Epidemiologic patterns
of Hodgkin’s disease. Int J Cancer 8(2):192–201
Clavel J, Steliarova-Foucher E, Berger C, Danon
S, Valerianova Z (2006) Hodgkin’s disease incidence and survival in European children and adolescents (1978–1997): report from the automated
Cancer information system project. Eur J Cancer
42(13):2037–2049
Hjalgrim LL, Rostgaard K, Engholm G et al (2016)
Aetiologic heterogeneity in pediatric Hodgkin lymphoma? Evidence from the Nordic countries, 1978–
2010. Acta Oncol (Stockholm) 55(1):85–90
Clarke CA, Glaser SL, Keegan THM, Stroup A
(2005) Neighborhood socioeconomic status and
Hodgkin’s lymphoma incidence in California.
Cancer Epidemiol Biomark Prev 14(6):1441–1447
Glaser SL, Swartz WG (1990) Time trends in
Hodgkin’s disease incidence. The role of diagnostic
accuracy. Cancer 66(10):2196–2204
Hjalgrim H, Askling J, Pukkala E et al (2001)
Incidence of Hodgkin’s disease in Nordic countries.
Lancet 358:297–298
van Leeuwen MT, Turner JJ, Joske DJ et al (2014)
Lymphoid neoplasm incidence by WHO subtype in
Australia 1982–2006. Int J Cancer 135(9):2146–2156
27. Hjalgrim H, Seow A, Rostgaard K, Friborg J (2008)
Changing patterns of Hodgkin lymphoma incidence
in Singapore. Int J Cancer 123(3):716–719
28. Zhu C, Bassig BA, Shi K et al (2014) Different time
trends by gender for the incidence of Hodgkin’s
lymphoma among young adults in the USA: a
birth cohort phenomenon. Cancer Causes Control
25(8):923–931
29. Aben KK, van Gaal C, van Gils NA, van der Graaf
WT, Zielhuis GA (2012) Cancer in adolescents and
young adults (15–29 years): a population-based
study in the Netherlands 1989–2009. Acta Oncol
(Stockholm) 51(7):922–933
30. Glaser SL, Clarke CA, Keegan THM, Chang ET,
Weisenburger DD (2015) Time trends in rates of
Hodgkin lymphoma histologic subtypes: true incidence changes or evolving diagnostic practice?
Cancer Epidemiol Biomark Prev 24(10):1474–1488
31. Poppema S, Van Imhoff G, Torensma R, Smit J
(1985) Lymphadenopathy morphologically consistent with Hodgkin’s disease associated with Epstein-­
Barr virus infection. Am J Clin Pathol 84(3):385–390
32. Lee J-H, Kim Y, Choi J-W, Kim Y-S (2014) Prevalence
and prognostic significance of Epstein-Barr virus
infection in classical Hodgkin’s lymphoma: a meta-­
analysis. Arch Med Res 45(5):417–431
33. Hjalgrim H (2012) On the aetiology of Hodgkin
lymphoma. Dan Med J 59(7):B4485
34. Razis DV, Diamond HD, Craver LF (1959)
Hodgkin’s disease associated with other malignant
tumors and certain non-neoplastic diseases. Am J
Med Sci 238:327–335
35. Kharazmi E, Fallah M, Pukkala E et al (2015) Risk
of familial classical Hodgkin lymphoma by relationship, histology, age, and sex: a joint study from five
Nordic countries. Blood 126(17):1990–1995
36. Mack TM, Cozen W, Shibata DK et al (1995)
Concordance for Hodgkin’s disease in identical twins suggesting genetic susceptibility to the
young-adult form of the disease. N Engl J Med
332(7):413–418
37. Goldin LR, Pfeiffer RM, Gridley G et al (2004)
Familial aggregation of Hodgkin lymphoma and
related tumors. Cancer 100(9):1902–1908
38. Goldin LR, Björkholm M, Kristinsson SY, Turesson
I, Landgren O (2009) Highly increased familial risks
for specific lymphoma subtypes. Br J Haematol
146(1):91–94
39. Amiel JL, Terasaki P (1967) Study of leucocyte
phenotypes in Hodgkin’s disease. In: Curtoni ES,
Mattiuz PL, Tosi MR (eds) Histocompatibility testing. Munksgaaid, Copenhagen, pp 79–81
40. Hors J, Dausset J (1983) HLA and susceptibility to
Hodgkin’s disease. Immunol Rev 70:167–192
41. Sud A, Thomsen H, Orlando G et al (2018) Genome-­
wide association study implicates immune dysfunction in the development of Hodgkin lymphoma.
Blood 132(19):2040–2052
42. Urayama KY, Jarrett RF, Hjalgrim H et al (2012)
Genome-wide association study of classical Hodgkin
1
Epidemiology of Hodgkin Lymphoma
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
lymphoma and Epstein-Barr virus status-defined
subgroups. J Natl Cancer Inst 104(3):240–253
Delahaye-Sourdeix M, Urayama KY, Gaborieau V
et al (2015) A novel risk locus at 6p21.3 for Epstein-­
Barr virus-positive Hodgkin lymphoma. Cancer
Epidemiol Biomark Prev 24(12):1838–1843
Sud A, Thomsen H, Law PJ et al (2017) Genome-­
wide association study of classical Hodgkin
lymphoma identifies key regulators of disease susceptibility. Nat Commun 8(1):1–11
Newell GR (1970) Etiology of multiple sclerosis and
Hodgkin’s disease. Am J Epidemiol 91(2):119–122
Gutensohn N, Cole P (1977) Epidemiology of
Hodgkin’s disease in the young. Int J Cancer
19:595–604
Gutensohn N, Cole P (1981) Childhood social
environment and Hodgkin’s disease. N Engl J Med
304(3):135–140
Mack TM, Norman JE, Rappaport E, Cozen W
(2015) Childhood determination of Hodgkin lymphoma among U.S. servicemen. Cancer Epidemiol
Biomark Prev 24(11):1707–1715
Hjalgrim H, Ekström Smedby K, Rostgaard K et al
(2007) Infectious mononucleosis, childhood social
environment, and risk of Hodgkin lymphoma.
Cancer Res 67(5):2382–2388
Glaser SL, Keegan THM, Clarke CA et al (2005)
Exposure to childhood infections and risk of
Epstein-Barr virus-defined Hodgkin’s lymphoma in
women. Int J Cancer 115(4):599–605
Glaser SL, Clarke CA, Nugent RA, Stearns CB,
Dorfman RF (2002) Social class and risk of
Hodgkin’s disease in young-adult women in 1988–
94. Int J Cancer 98(1):110–117
Chang ET, Zheng T, Weir EG et al (2004)
Childhood social environment and Hodgkin’s lymphoma: new findings from a population-based casecontrol study. Cancer Epidemiol Biomark Prev
13(8):1361–1370
Crump C, Sundquist K, Sieh W, Winkleby MA,
Sundquist J (2012) Perinatal and family risk factors
for Hodgkin lymphoma in childhood through young
adulthood. Am J Epidemiol 176(12):1147–1158
Triebwasser C, Wang R, DeWan AT et al (2016)
Birth weight and risk of paediatric Hodgkin lymphoma: findings from a population-based record
linkage study in California. Eur J Cancer 69:19–27
Westergaard T, Melbye M, Pedersen JB et al (1997)
Birth order, sibship size and risk of Hodgkin’s disease in children and young adults: a population-­
based study of 31 million person-years. Int J Cancer
72(6):977–981
Chang ET, Montgomery SM, Richiardi L et al (2004)
Number of siblings and risk of Hodgkin’s lymphoma.
Cancer Epidemiol Biomark Prev 13(7):1236–1243
Von Behren J, Spector LG, Mueller BA et al
(2011) Birth order and risk of childhood cancer: a
pooled analysis from five US states. Int J Cancer
128(11):2709–2716
21
58. Linabery AM, Erhardt EB, Fonstad RK et al (2014)
Infectious, autoimmune and allergic diseases and
risk of Hodgkin lymphoma in children and adolescents: a Children’s oncology group study. Int J
Cancer 135(6):1454–1469
59. Isager H, Andersen E (1978) Pre-morbid factors in
Hodgkin’s disease. I. Birth weight and growth pattern from 8 to 14 years of age. Scand J Haematol
21(3):250–255
60. Keegan THM, Glaser SL, Clarke CA et al (2006)
Body size, physical activity, and risk of Hodgkin’s
lymphoma in women. Cancer Epidemiol Biomark
Prev 15(6):1095–1101
61. Larsson SC, Wolk A (2011) Body mass index and
risk of non-Hodgkin’s and Hodgkin’s lymphoma: a
meta-analysis of prospective studies. Eur J Cancer
47(16):2422–2430
62. Engeland A, Tretli S, Hansen S, Bjørge T (2007)
Height and body mass index and risk of lymphohematopoietic malignancies in two million Norwegian
men and women. Am J Epidemiol 165(1):44–52
63. Murphy F, Kroll ME, Pirie K et al (2013) Body size in
relation to incidence of subtypes of ­haematological
malignancy in the prospective million women study.
Br J Cancer 108(11):2390–2398
64. Epstein MA, Achong BG, Barr YM (1964) Virus
particles in cultured lymphoblasts from Burkitt’s
lymphoma. Lancet 1:702–703
65. Henle G, Henle W, Diehl V (1968) Relation of
Burkitt’s tumor-associated herpes-type virus to
infectious mononucleosis. Proc Natl Acad Sci U S A
59(1):94–101
66. Odumade OA, Hogquist KA, Balfour HH (2011)
Progress and problems in understanding and managing primary Epstein-Barr virus infections. Clin
Microbiol Rev 24(1):193–209
67. Hjalgrim H, Askling J, Sørensen P et al (2000)
Risk of Hodgkin’s disease and other cancers after
infectious mononucleosis. J Natl Cancer Inst
92(18):1522–1528
68. Hjalgrim H, Askling J, Rostgaard K et al
(2003) Characteristics of Hodgkin’s lymphoma
after infectious mononucleosis. N Engl J Med
349(14):1324–1332
69. Alexander FE, Jarrett RF, Lawrence D et al (2000)
Risk factors for Hodgkin’s disease by Epstein-Barr
virus (EBV) status: prior infection by EBV and other
agents. Br J Cancer 82(5):1117–1121
70. Alexander FE, Lawrence DJ, Freeland J et al
(2003) An epidemiologic study of index and family infectious mononucleosis and adult Hodgkin’s
disease (HD): evidence for a specific association
with EBV +ve HD in young adults. Int J Cancer
107(2):298–302
71. IARC (1997) IARC monographs on the evaluation
of carcinogenic risks to humans: tobacco smoke and
involuntary smoking. IARC, Lyon, p 83
72. Coghill AE, Hildesheim A (2014) Epstein-Barr
virus antibodies and the risk of associated malig-
H. Hjalgrim and R. F. Jarrett
22
73.
74.
75.
76.
77.
78.
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
nancies: review of the literature. Am J Epidemiol
180(7):687–695
Mueller N, Evans A, Harris NL et al (1989)
Hodgkin’s disease and Epstein-Barr virus. Altered
antibody pattern before diagnosis. N Engl J Med
320(11):689–695
Levin LI, Chang ET, Ambinder RF et al (2012)
Atypical prediagnosis Epstein-Barr virus serology
restricted to EBV-positive Hodgkin lymphoma.
Blood 120(18):3750–3755
Glaser SL, Lin RJ, Stewart SL et al (1997) Epstein-­
Barr virus-associated Hodgkin’s disease: epidemiologic characteristics in international data. Int J
Cancer 70(4):375–382
Viruses JRF (2002) Hodgkin’s lymphoma. Ann
Oncol 13(Suppl 1):23–29
Monnereau A, Orsi L, Troussard X et al (2007)
History of infections and vaccinations and risk of
lymphoid neoplasms: does influenza immunization
reduce the risk? Leukemia 21(9):2075–2079
Montella M, Maso LD, Crispo A et al (2006) Do
childhood diseases affect NHL and HL risk? A case-­
control study from northern and southern Italy. Leuk
Res 30(8):917–922
Cozen W, Hamilton AS, Zhao P et al (2009) A
protective role for early oral exposures in the etiology of young adult Hodgkin lymphoma. Blood
114(19):4014–4020
Engels EA, Hildesheim A (2018) Immunologic factors. In: Schottenfeld and Fraumeni cancer epidemiology and prevention. Oxford University Press, Oxford
Frisch M, Biggar RJ, Engels EA et al (2001)
Association of cancer with AIDS-related immunosuppression in adults. JAMA 285(13):1736
Shepherd L, Ryom L, Law M et al (2018)
Differences in virological and immunological risk
factors for non-Hodgkin and Hodgkin lymphoma. J
Natl Cancer Inst 110(6):598–607
Grulich AE, Vajdic CM (2015) The epidemiology
of cancers in human immunodeficiency virus infection and after organ transplantation. Semin Oncol
42(2):247–257
Landgren O, Caporaso NE (2007) New aspects in
descriptive, etiologic, and molecular epidemiology
of Hodgkin’s lymphoma. Hematol Oncol Clin North
Am 21(5):825–840
Fallah M, Liu X, Ji J et al (2014) Hodgkin lymphoma after autoimmune diseases by age at diagnosis and histological subtype. J Eur Soc Med Oncol
25(7):1397–1404
Hemminki K, Försti A, Sundquist K, Sundquist J,
Li X (2017) Familial associations of lymphoma and
myeloma with autoimmune diseases. Blood Cancer
J 7(1):e515–e515
Khankhanian P, Cozen W, Himmelstein DS et al (2016)
Meta-analysis of genome-wide association studies
reveals genetic overlap between Hodgkin lymphoma
and multiple sclerosis. Int J Epidemiol 45(3):728–740
Montgomery S, Hajiebrahimi M, Burkill S et al
(2016) Multiple sclerosis and risk of young-adult-­
89.
90.
91.
92.
93.
94.
95.
96.
97.
98.
99.
100.
101.
102.
103.
onset Hodgkin lymphoma. Neurol Neuroimmunol
Neuroinflamm 3(3):e227
Hjalgrim H, Rasmussen S, Rostgaard K et al
(2004) Familial clustering of Hodgkin lymphoma and multiple sclerosis. J Natl Cancer Inst
96(10):780–784
Hollander P, Rostgaard K, Smedby KE et al (2015)
Autoimmune and atopic disorders and risk of
classical Hodgkin lymphoma. Am J Epidemiol
182(7):624–632
Martínez-Maza O, Moreno AD, Cozen W (2010)
Epidemiological evidence: IgE, allergies, and hematopoietic malignancies. Cancer IgE 2010:79–136
Levin LI, Breen EC, Birmann BM et al (2017)
Elevated serum levels of sCD30 and IL6 and
detectable IL10 precede classical Hodgkin lymphoma diagnosis. Cancer Epidemiol Biomark Prev
26(7):1114–1123
Chang ET, Zheng T, Weir EG et al (2004) Aspirin
and the risk of Hodgkin’s lymphoma in a population-­
based case-control study. J Natl Cancer Inst
96(4):305–315
Chang ET, Frøslev T, Sørensen HT, Pedersen L (2011)
A nationwide study of aspirin, other non-steroidal
anti-inflammatory drugs, and Hodgkin lymphoma
risk in Denmark. Br J Cancer 105(11):1776–1782
Mondul AM, Weinstein SJ, Layne TM, Albanes D
(2017) Vitamin D and cancer risk and mortality:
state of the science, gaps, and challenges. Epidemiol
Rev 39(1):28–48
Bowen EM, Pfeiffer RM, Linet MS et al (2016)
Relationship between ambient ultraviolet radiation
and Hodgkin lymphoma subtypes in the United
States. Br J Cancer 114(7):826–831
van Leeuwen MT, Turner JJ, Falster MO et al (2013)
Latitude gradients for lymphoid neoplasm subtypes
in Australia support an association with ultraviolet
radiation exposure. Int J Cancer 133(4):944–951
Monnereau A, Glaser SL, Schupp CW et al
(2013) Exposure to UV radiation and risk of
Hodgkin lymphoma: a pooled analysis. Blood
122(20):3492–3499
Sopori M (2002) Effects of cigarette smoke
on the immune system. Nat Rev Immunol
2(May):372–377
Sergentanis TN, Kanavidis P, Michelakos T, Petridou
ET (2013) Cigarette smoking and risk of lymphoma
in adults. Eur J Cancer Prev 22(2):131–150
Castillo JJ, Dalia S, Shum H (2011) Meta-analysis
of the association between cigarette smoking and
incidence of Hodgkin’s lymphoma. J Clin Oncol
29(29):3900–3906
Kamper-Jørgensen M, Rostgaard K, Glaser SL et al
(2013) Cigarette smoking and risk of Hodgkin lymphoma and its subtypes: a pooled analysis from the
international lymphoma epidemiology consortium
(InterLymph). Ann Oncol 24(9):2245–2255
Klatsky AL, Li Y, Baer D et al (2009) Alcohol consumption and risk of hematologic malignancies. Ann
Epidemiol 19(10):746–753
1
Epidemiology of Hodgkin Lymphoma
104. Lim U, Morton LM, Subar AF et al (2007) Alcohol,
smoking, and body size in relation to incident
Hodgkin’s and non-Hodgkin’s lymphoma risk. Am
J Epidemiol 166(6):697–708
105. Bernard SM, Cartwright RA, Darwin CM et al
(1987) Hodgkin’s disease: case control epidemiological study in Yorkshire. Br J Cancer
55(1):85–90
106. Besson H, Brennan P, Becker N et al (2006) Tobacco
smoking, alcohol drinking and Hodgkin’s lymphoma: a European multi-Centre case-control study
(EPILYMPH). Br J Cancer 95(3):378–384
107. Monnereau A, Orsi L, Troussard X et al (2008)
Cigarette smoking, alcohol drinking, and risk
of lymphoid neoplasms: results of a French
case-­
control study. Cancer Causes Control
19(10):1147–1160
108. Nieters A, Deeg E, Becker N (2006) Tobacco
and alcohol consumption and risk of lymphoma:
results of a population-based case-control study in
Germany. Int J Cancer 118(2):422–430
109. Kanda J, Matsuo K, Inoue M et al (2010) Association
of alcohol intake with the risk of malignant lym-
23
110.
111.
112.
113.
114.
phoma and plasma cell myeloma in Japanese: a
population-based cohort study (Japan public health
center-based prospective study). Cancer Epidemiol
Biomark Prev 19(2):429–434
Willett EV, O’Connor S, Smith AG, Roman E (2007)
Does smoking or alcohol modify the risk of Epstein-­
Barr virus-positive or -negative Hodgkin lymphoma?
Epidemiology 18(1):130–136
Gorini G, Stagnaro E, Fontana V et al (2007) Alcohol
consumption and risk of Hodgkin’s lymphoma and
multiple myeloma: a multicentre case-control study.
Ann Oncol 18(1):143–148
Kroll ME, Murphy F, Pirie K et al (2012) Alcohol
drinking, tobacco smoking and subtypes of haematological malignancy in the UK million women study.
Br J Cancer 107(5):879–887
Psaltopoulou T, Sergentanis TN, Ntanasis-­
Stathopoulos I et al (2018) Alcohol consumption and
risk of hematological malignancies: a meta-analysis
of prospective studies. Int J Cancer 143(3):486–495
Stein RS, Morgan DS (2004) Hodgkin disease. In:
Wintrobe’s clinical hematology. Lippincott Williams
& Wilkins, Philadelphia, PA
2
The Role of Viruses in the Genesis
of Hodgkin Lymphoma
Ruth F. Jarrett, Henrik Hjalgrim,
and Paul G. Murray
Contents
2.1
Introduction 26
2.2
Hodgkin Lymphoma and Epstein-Barr Virus 2.2.1 E
pstein-Barr Virus and the Pathogenesis of Hodgkin Lymphoma 2.2.2 Risk Factors for Epstein-Barr Virus-Associated Hodgkin Lymphoma 2.2.3 Epstein-Barr Virus and Hodgkin Lymphoma: A Causative Association? 2.2.4 Epstein-Barr Virus and the Clinicopathological Features of Hodgkin
Lymphoma 26
27
29
32
2.3
Epstein-Barr Virus-Negative Hodgkin Lymphoma Cases 2.3.1 H
odgkin Lymphoma and Herpesviruses Other Than Epstein-Barr Virus 2.3.2 Polyomaviruses and Hodgkin Lymphoma 2.3.3 Measles Virus and Hodgkin Lymphoma 2.3.4 The Virome, Anelloviruses, and Hodgkin Lymphoma 34
34
36
37
37
2.4
Conclusions 33
38
References 38
Abbreviations
R. F. Jarrett (*)
MRC-University of Glasgow Centre for Virus
Research, Glasgow, UK
e-mail: ruth.jarrett@glasgow.ac.uk
BARTs
H. Hjalgrim
Department of Epidemiology Research, Statens
Serum Institut, Copenhagen, Denmark
cHL
DDR1
EBER
EBNA
EBV
HHV
HL
HLA
HPyV
HRS
Department of Haematology, Copenhagen University
Hospital Rigshospitalet, Copenhagen, Denmark
e-mail: HHJ@ssi.dk
P. G. Murray
Bernal Institute, Limerick, Ireland
Institute of Immunology and Immunotherapy,
University of Birmingham, Birmingham, UK
e-mail: Paul.Murray@ul.ie
BHRF1
BamHI fragment A rightward
transcripts
BamHI-H rightward open reading
frame 1
Classic Hodgkin lymphoma
Discoidin domain receptor 1
EBV-encoded small RNAs
EBV nuclear antigen
Epstein-Barr virus
Human herpesvirus
Hodgkin lymphoma
Human leukocyte antigen
Human polyomavirus
Hodgkin and Reed-Sternberg
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_2
25
R. F. Jarrett et al.
26
IHC
LMP
MCHL
Immunohistochemistry
Latent membrane protein
Mixed
cellularity
Hodgkin
lymphoma
Merkel cell polyomavirus
MicroRNAs
Measles virus
Nodular
sclerosis
Hodgkin
lymphoma
Open reading frame
Polyomavirus
Single nucleotide polymorphism
Trichodysplasia
spinulosa
polyomavirus
Torque teno midi virus
Torque teno mini virus
Torque teno virus
2.2
Hodgkin Lymphoma
and Epstein-Barr Virus
EBV is a herpesvirus with a worldwide distribution [10–13]. Over 90% of healthy adults are
infected by EBV, and, following primary infection, the virus establishes a persistent infection
with a reservoir in memory B-cells [14]. Although
EBV is an extremely efficient transforming
ORF
agent, the virus is kept under tight control by cell-­
PyV
mediated immune responses, and both primary
SNP
and persistent infections are usually asymptomTSPyV
atic [10, 15].
EBV infection can be lytic or latent. Lytic
TTMDV
infection is associated with expression of a large
TTMV
number of viral genes, production of progeny
TTV
virus, and death of the infected cell; in contrast,
latent infection is associated with expression of a
small number of EBV genes, persistent infection,
and growth transformation [10]. In B-cells trans2.1
Introduction
formed by EBV in vitro, six EBV nuclear antigens (EBNA1, 2, 3A, 3B, 3C, and LP, also called
Hodgkin lymphoma (HL) is a heterogeneous con- EBNA1–6) and three latent membrane proteins
dition. Seminal papers published in 1957 and (LMP1, LMP2A, and LMP2B) are expressed
1966 suggested that HL in younger and older [10]. In addition, noncoding viral RNAs are tranadults had different etiologies and further sug- scribed in latently infected cells [16]. These
gested an infectious etiology for young adult HL include two small non-polyadenylated tran[1, 2]. Subsequent epidemiological studies pro- scripts, the EBERs, and over 44 viral microRNAs
vide broad support for these hypotheses [3, 4]. (miRNAs) located within introns of the BARTs
Data linking young adult HL with a high standard (BamHI fragment A rightward transcripts) or
of living in early childhood and lack of child-­child around the coding region of the BHRF1
contact suggest that delayed exposure to common (BamHI-H rightward open reading frame 1) gene
childhood infections may be involved in the etiol- [16–22]. Expression of the full set of latent genes
ogy of these cases [5, 6]. There is now compelling is known as latency III [10, 13]. EBV gene
evidence that a proportion of cases of HL are expression in EBV-positive lymphomas occurassociated with the Epstein-Barr virus (EBV). ring in the context of immunosuppression freParadoxically, older adult and childhood cases of quently follows this pattern, but more restricted
HL are more likely to be EBV-associated than patterns of EBV gene expression are observed in
young adult cases [7–9]. In this article, we review other malignancies, including cHL [10, 12, 13].
studies on viral involvement in HL with a focus The EBNA3 family proteins are immunodomion classic HL (cHL), since nodular lymphocyte-­ nant, and the other latent antigens elicit only subpredominant HL is considered a separate disease dominant or weak cell-mediated immune
entity. The association with EBV will be dis- responses [23, 24]. The pattern of gene exprescussed with an emphasis on findings that support sion in EBV-associated malignancies most proba causal role for EBV in this malignancy. Studies ably depends on both the lineage and stage of
investigating the direct involvement of other differentiation of the infected tumor cells and the
exogenous viruses will be summarized.
host EBV-specific immune response.
MCPyV
miRNAs
MV
NSHL
2
The Role of Viruses in the Genesis of Hodgkin Lymphoma
In EBV-associated cHL (also referred to as
EBV-positive cHL), all of the tumor cells, the
Hodgkin and Reed-Sternberg (HRS) cells, are
infected by EBV [25–27]. The EBV infection
within tumors is also clonal suggesting that all of
the tumor cells are derived from a single infected
cell [28, 29]. The HRS cells express EBNA1,
LMP1, LMP2A, and 2B, but the remaining
EBNAs are downregulated (Fig. 2.1); the noncoding EBER RNAs and BART miRNAs are also
expressed [25, 26, 30–33]. This pattern of gene
expression is referred to as latency type II [10].
EBV infection of HRS cells can be readily
demonstrated in sections of routinely fixed,
paraffin-­embedded material using either EBER
in situ hybridization or LMP1 immunohistochemistry (IHC) (Fig. 2.1) [25, 26]. Reagents for
both assays are commercially available.
2.2.1
27
35]. A pathognomonic feature of these cells is the
global suppression of B-cell signature genes and
inappropriate expression of genes associated with
other hematopoietic lineages [36, 37]. Importantly,
HRS cells do not express B-cell receptors (BCRs).
Survival of germinal center B-cells normally
requires signaling through both BCRs and CD40;
HRS cells must, therefore, have acquired a nonphysiological survival mechanism(s). Functional
studies of EBV, and LMP1 and LMP2A, support a
role for the virus in HRS cell survival, transcriptional reprogramming, and immune evasion, as
summarized below (Fig. 2.2).
Epstein-Barr Virus
and the Pathogenesis
of Hodgkin Lymphoma
The molecular pathogenesis of cHL and the origin
of the HRS cell are described in detail in Chap. 3.
Briefly, HRS cells have clonally rearranged immunoglobulin genes with evidence of somatic hypermutation, indicating a derivation from B-cells that
have participated in a germinal center reaction [34,
Fig. 2.2 EBV EBER in situ hybridization staining of
EBV-positive Hodgkin and Reed-Sternberg cells. The
characteristic staining pattern is observed in the nuclei of
Hodgkin and Reed-Sternberg cells
LMP2A
LMP1
C-terminus activates signalling pathways
e.g. NF-ĸB, JAK/STAT, PI3-K/Akt
Induces chemokine/cytokine secretion
that promotes tumour microenvironment
Fig. 2.1 The latent membrane proteins of EBV contribute to the pathogenesis of classic Hodgkin lymphoma.
Schematic diagram of LMP1 (left) and LMP2A (right)
proteins in the cell membrane (gray bar). Both are transmembrane proteins that signal constitutively through the
C-terminus in the case of LMP1 and the N-terminus in the
LMP1 LMP2
N-terminus activates signalling pathways
e.g. PI3-K/Akt
Contributes to survival of BCR-negative
B-cells
case of LMP2A. The photomicrograph in the center shows
the co-expression of LMP1 (red) and LMP2A (green) in
the same Hodgkin and Reed-Sternberg cell in a tissue section of classic Hodgkin lymphoma. The nucleus of the
Hodgkin and Reed-Sternberg cell stained blue with DAPI
is arrowed
28
In 2005, three independent groups published
data showing that germinal center B-cells lacking
BCRs could survive and be immortalized by
EBV [38–40]. In elegant experiments, Mancao
and Hammerschmidt later showed that this survival function was dependent on LMP2A expression [41]. A series of in vivo and in vitro studies
from the Longnecker laboratory further defined
LMP2A function and showed that this viral protein could mimic an activated BCR and provide a
survival signal to BCR-negative B-cells [42–44].
LMP2A expression in B-cells also results in
downregulation of B-cell-specific genes and
induction of genes associated with proliferation
and inhibition of apoptosis, a gene expression
profile similar to that seen in cHL-derived cell
lines [45]. Constitutive activation of Notch1 by
LMP2A, and subsequent inhibition of E2A and
downregulation of EBF, two transcription factors
that regulate B-cell development, appears to be
involved in both survival signaling and transcriptional regulation [44]. Although these data suggest a role for LMP2A in the survival and
reprogramming of HRS cells, many of the intracellular molecules involved in BCR signaling are
downregulated in HRS cells, and therefore the
precise contribution of LMP2A in cHL is not
clear.
CD40 signaling plays a critical role in the positive selection of germinal center B-cells expressing high-affinity immunoglobulin and their
subsequent exit from the germinal center [46].
EBV LMP1 is an integral membrane protein
which interacts with several signal transduction
pathways to activate NF-κB, Jun N-terminal
kinase (JNK), and p38 mitogen-activated protein
kinase [47–51]. In this way, LMP1 mimics a constitutively active CD40 molecule, although it provides a more potent and sustained signal. Many
of the genes that are transcriptionally regulated
by LMP1 in germinal center B-cells are also
CD40 and NF-κB targets [52]. Activation of the
NF-κB pathway, which is a feature of both EBV-­
positive and EBV-negative HRS cells, leads to
upregulation of anti-apoptotic genes and is
thought to play a key role in HRS cell survival
[53–55]. LMP1 expression in germinal center
B-cells also leads to increased expression of Id2,
R. F. Jarrett et al.
an inhibitor of the E2A transcription factor mentioned above, and repression of B-cell signature
genes [52]; therefore, LMP1 may also contribute
to transcriptional reprogramming.
The EBV genome is maintained as an episome
in infected cells; i.e., it does not normally integrate. The EBNA1 protein is responsible for
maintenance of the genome in episomal form,
genome replication, and genome partitioning
during mitosis [10, 56]. EBNA1 can also influence both viral and cellular gene expression and
appears to confer a B-cell survival advantage,
although the impact of EBNA1 on oncogenesis
in vivo is controversial [10, 57–60]. Interestingly,
in the context of cHL, overexpression of EBNA1
in vitro leads to the appearance of multinucleated
cells [57].
The EBV EBER RNAs are two small, non-­
polyadenylated RNA pol III transcripts that are
stably expressed in the nuclei of all latently
infected cells, including HRS cells. The precise
function of the EBERs remains unclear, and,
although not essential for transformation, expression of these small RNAs is required for efficient
EBV-induced B-cell growth and transformation
[16, 61–63].
EBV-encoded miRNAs were identified first in
2004, and their important role in EBV biology
and oncogenesis is an area of intense study [16,
17, 22, 64, 65]. Functional analysis of the BHRF1
and BART miRNAs suggests roles in evading the
immune response, promoting cell survival and
proliferation, inhibiting viral reactivation, and
fine-tuning gene expression [16, 22, 64, 65].
EBV-associated malignancies, including cHL,
express BART miRNAs, but the BHRF1 miRNAs, which are associated with latency type III,
are not expressed [33]. In vitro studies of knockout viruses lacking some or all of the miRNAs
suggest that they have an important role in the
initial stages of B-cell transformation by EBV;
BHRF1 miRNAs play the predominant role with
some contribution from BART miRNAs at low
multiplicity of infection [22]. In contrast, in vivo
studies in a murine huNSG model suggest that
the main function of these miRNAs is to attenuate the antiviral T-cell-mediated immune
response, leading to increased numbers of EBV-­
2
The Role of Viruses in the Genesis of Hodgkin Lymphoma
infected B-cells at later time points [66]. Again,
these effects appear to be mediated by the BHRF1
miRNAs, as viruses deficient in only the BART
miRNAs produced similar results to wild-type
virus in this model system [66]. Ross et al.
reported that miRNA BART11-5p downregulates
the B-cell transcription factor EBF1, suggesting a
plausible role for this miRNA in cHL [67]. EBV
also regulates the expression of host miRNAs;
infection of primary B-cells leads to a conspicuous downregulation of many miRNAs with the
notable exception of mIR-155, which is highly
expressed by both EBV-positive and EBV-­
negative HRS cells [68, 69]. Analysis of host
miRNAs in cHL is described in more detail in
Chap. 4, but it has been reported that EBV status
of tumors is associated with differences in expression pattern [70].
While most studies have investigated the
effects of the latent genes in isolation, there is
evidence that co-expression of the EBV latent
genes is important. For example, it has been
shown that LMP1 is transforming when
expressed alone in transgenic mouse B-cells
[71]. However, when LMP2A is expressed
together with LMP1, the resulting mouse B-cells
are normal [71]. Comparison of LMP1 and
LMP2A in B-cells confirms they have both synergistic and counteracting transcriptional effects
[72]. Furthermore, in another study it was shown
that LMP1 and LMP2A co-expression in mouse
B-cells resulted in tumors, but only if the animals were immunosuppressed suggesting that
the combined expression of these latent genes is
immunogenic in vivo [73].
There is evidence that the tumor microenvironment in EBV-positive and EBV-negative cHL
is different. Thus, the T-helper cells present in
EBV-positive cHL are enriched for functional
Th1 cells [74]. EBV-positive cHL is also preferentially infiltrated by regulatory Type 1 cells
(Tr1), which express ITGA2, ITGB2, and LAG3
and secrete IL-10 [75]. This Th1-biased infiltrate
is consistent with previous reports of higher
numbers of activated CD8+ T-cells in EBVpositive cHL [76] and is also associated with the
presence of predominantly M1-polarized macrophages [77]. There is evidence that the EBV
29
latent genes are responsible, at least in part, for
the recruitment and modification of this tumor
microenvironment. LMP1, in particular, has been
shown to induce expression of many of the chemokines and cytokines secreted by EBV-infected
HRS cells [78, 79]. The cHL tumor microenvironment also contributes to the suppression of
host anti-­
EBV-­
specific immunity. Thus, while
LMP1 and LMP2A proteins are targets of CD8+
cytotoxic T lymphocytes, it is clear that immune
effectors present in the tumor tissues of EBVpositive cHL cannot kill the virus-infected cells
[80, 81]. LMP1 probably also contributes to the
suppression of EBV-specific immunity through
its ability to induce expression of the immunosuppressive cytokines, including IL-10, and
upregulate the immune checkpoint ligand, PD-L1
[82, 83]. LMP1 also upregulates the collagen
receptor, discoidin domain receptor 1 (DDR1), a
receptor tyrosine kinase expressed by HRS cells
[84]. Engagement of DDR1 by collagen leads to
the increased survival of lymphoma cells, thus
providing a link between the expression of LMP1
and pro-survival signaling from the tumor
microenvironment.
2.2.2
isk Factors for Epstein-Barr
R
Virus-Associated Hodgkin
Lymphoma
It is clear that EBV is associated with only a proportion of cHL cases. In industrialized countries
around one third of cases are EBV-associated,
whereas in Africa and Central and South America,
this proportion is significantly higher [8, 9, 85, 86].
EBV-associated cHL cases are not randomly distributed among all cHL cases, and the demographic
features and risk factors for the development of
EBV-positive and EBV-negative cHL show distinctive features [8, 9, 86]. Childhood (<10 years) and
older adult (50+ years) cases are more likely to
be EBV-associated than young adult cases
(15–34 years) [7, 8, 86]. Among EBV-­associated
cases, males outnumber females by approximately
2:1, whereas males and females are more evenly
represented among EBV-­negative cases [9, 86]. In
developing countries, childhood cHL is more com-
R. F. Jarrett et al.
30
mon than in industrialized countries, resulting in a
higher proportion of EBV-associated cases [9, 86,
87]. Material deprivation is associated with an
increased proportion of EBV-positive childhood
cHL cases in industrialized countries, and there is
some evidence that this also holds for older adult
cases [85, 88].
EBV infection usually occurs in childhood,
and in many parts of the world, there is almost
universal infection by the age of 5 years [11, 89].
If infection is delayed until adolescence, as is
increasingly observed in industrialized countries,
primary EBV infection manifests as infectious
mononucleosis in around 25% of individuals
[90]. Infectious mononucleosis has been associated with an increased risk of EBV-associated
cHL in some, although not all, studies [6, 91–93].
The increased risk appears relatively short-lived
with a median time interval between infectious
mononucleosis and cHL of approximately
3–4 years (see Chap. 1: Epidemiology of Hodgkin
Lymphoma) [92, 93]. Thus, in both developing
and developed countries, there appears to be a
period following primary EBV infection, probably lasting several years, in which risk of EBV-­
associated cHL is increased. cHL occurring in
the context of immunosuppression is almost
always EBV-associated (see Chap. 1:
Epidemiology of Hodgkin Lymphoma) [94, 95],
and it is likely that the increased incidence of
EBV-associated cHL that occurs in older adults is
related to immune senescence. Based on these
data, we have proposed an extension of
MacMahon’s model of HL that divides cHL into
four subgroups on the basis of tumor EBV status,
age at diagnosis, and age at infection by EBV
(Fig. 2.3) [2, 96].
Recent data also suggest that humoral and
cell-mediated responses to EBV modulate risk of
EBV-associated cHL. Levin and colleagues
examined anti-EBV antibody profiles in serum
samples from military personnel (mainly young
men) that had been collected several years before
the diagnosis of cHL [97]. Individuals who subsequently developed EBV-positive, but not EBV-­
negative, cHL were more likely to have elevated
Incidence
4
3
1
2
10
25
50
Age (years)
Fig. 2.3 The four-disease model of classic Hodgkin lymphoma. This model divides classic Hodgkin lymphoma
into four subgroups based on EBV tumor status, age at
diagnosis, and age at EBV infection. Three groups of
EBV-associated disease are recognized: (1) a childhood
disease, usually occurring below the age of 10 years,
which is commoner in developing countries; (2) a disease,
most commonly seen in young adults, which occurs following infectious mononucleosis; and (3) a disease associated with poor control of EBV infection, which is
typified by the older adult cases but can occur at other
ages, particularly in the context of immunosuppression.
(4) Superimposed on these is a single group of EBV-­
negative classic Hodgkin lymphoma cases, which account
for the young adult age-specific incidence peak seen in
industrialized countries. The relative incidence of each of
these four disease subgroups will determine the overall
shape of the age-specific incidence curve in any geographical locale
2
The Role of Viruses in the Genesis of Hodgkin Lymphoma
antibody titers to EBV viral capsid and early antigens and an anti-EBNA-1/anti-EBNA2 antibody
ratio ≤1.0 when compared to controls. Decreased
anti-EBNA-1/anti-EBNA2 antibody ratios have
been previously associated with EBV-associated
cHL [98], and it has been suggested that a
ratio ≤ 1.0, which persists for more than 2 years
after infectious mononucleosis, indicates defective control of persistent EBV infection [99].
Variations in EBNA-1 titer are significantly associated with polymorphisms in the human leukocyte antigen (HLA) region [100], suggesting that
titers may, in part, be genetically determined and
relate to the findings described below.
Data from HLA association studies and
genome-wide association studies (GWAS) show
clear associations between cHL risk and both
HLA alleles and single nucleotide polymorphisms (SNPs) in this region. Although some
SNPs appear to be associated with all cHL, independent of EBV status, most HLA associations
differ between EBV-positive and EBV-negative
subgroups [101–108]. Both HLA class I and II
alleles are associated with EBV-positive cHL,
whereas EBV-negative cHL is largely associated
with class II alleles [102, 103, 105, 107]. Since
class I and II alleles present peptides from
31
pathogens to CD8- and CD4-positive T-cells,
respectively, this suggests that genetically determined differences in the cell-mediated response
to EBV influence disease risk. HLA-A∗01 is
associated with an increased risk of EBVassociated cHL, whereas HLA-A∗02, specifically A∗02:01, is associated with decreased risk
[102, 103]. Associations with these alleles are
independent, i.e., the increased risk associated
with A∗01 is not simply due to lack of A∗02, and
effects are dependent on the copy number of
each of the alleles [103] (Fig. 2.4). As a result,
there is an almost tenfold variation in odds of
EBV-­
associated cHL between HLA-A∗01
homozygotes and HLA-A∗02 homozygotes
[103]. More recent data suggest that B∗37:01 is
also ­associated with an increased risk of EBVpositive cHL [105, 107]. Class II alleles have
been less extensively studied, but Huang et al.
reported an increased frequency of DR10 alleles
in patients with EBV-positive cHL compared to
controls, and we have detected protective effects
of DRB1∗15:01 and DPB1∗01:01 [105, 107]. In
addition, the SNP rs6457715, which is located
close to the HLA-DPB1 gene, is strongly associated with EBV-positive but not EBV-negative
cHL [108].
Sex
Age group
A*01:01 add
A*02:01 add
IM
A*02:01 x IM interaction
0
2
4
6
8
Odds Ratio
Fig. 2.4 Risk factors for EBV-associated classic Hodgkin
lymphoma in adults. Forest plot showing odds ratios and
95% confidence intervals for development of EBVassociated Hodgkin lymphoma from a case series analysis
of HLA and non-HLA risk factors [103]. Increased risk is
associated with male sex, older age (age ≥50 years versus
15–34 years), possession of HLA-A∗01:01 alleles (add,
additive effect), and prior history of infectious mononucleosis (IM). Possession of HLA-A∗02:01 alleles is associated with decreased risk, and abrogation of the increased
risk associated with IM
R. F. Jarrett et al.
32
Cytotoxic T-cell responses, restricted through
HLA class I, are critical for the control of EBV
infection, and A∗02 is known to present a wide
range of peptides derived from EBV lytic and
latent antigens, including those expressed by
HRS cells [23, 24]. In contrast, there are no
well-­characterized A∗01-restricted EBV epitopes [24], and EBV-specific T-cell responses
restricted through A∗01:01 have not been
described [109]. The observed associations with
HLA-A, therefore, seem biologically plausible.
However, HLA-A∗01 is in strong linkage disequilibrium with HLA-B∗08, which is associated with immunodominant EBV-specific
cytotoxic T-cell responses, and yet there is no
protective effect associated with this allele
[107]. The biological basis underlying associations between HLA alleles and EBV-associated
cHL is therefore not clear. Further work is also
necessary to determine whether the critical
HLA-A-restricted
cell-­
mediated
immune
responses are directed toward EBV-infected
HRS cells or whether it is the control of persistent EBV infection, i.e., the host-­virus equilibrium, which is all-important. The increased risk
associated with individual class I alleles favors
the idea that failure to respond to a particular
protein, or very restricted group of proteins,
determines risk; this focusses attention on EBV
proteins expressed by HRS cells. Consistent
with this, no EBNA1, LMP1, or LMP2 epitopes
restricted by B∗37:01 have been identified
although a B∗37:01-restricted EBNA3C epitope
has been described [24].
As mentioned above, prior infectious mononucleosis has been associated with an increased
risk of EBV-positive cHL [91–93, 110].
Infectious mononucleosis has also been associated with the same genotypic markers (microsatellites and SNPs) in the HLA class I region as
EBV-positive cHL, albeit with lesser statistical
significance [111]. These data raised the possibility that the association between infectious
mononucleosis and EBV-associated cHL
resulted from shared genetic susceptibility.
However, HLA-A typing of over 700 cHL cases
with available self-­reported history of infectious
mononucleosis revealed that prior infectious
mononucleosis was independently associated
with EBV-associated cHL after adjusting for the
effects of HLA-A alleles [103]. In addition, a
statistically significant interaction between prior
infectious mononucleosis and HLA-A∗02 was
detected; the effect of this was to abrogate the
increased risk of EBV-associated cHL following
infectious mononucleosis in HLA-A∗02positive individuals [103]. These results suggest
that the increased risk of EBV-associated cHL
following infectious mononucleosis is modified
by the EBV-specific cytotoxic T-cell response
restricted through HLA-A∗02. Thus, it is possible that different HLA alleles exert their effects
at different stages in the natural history of EBVassociated cHL.
Associations with childhood cHL and infectious mononucleosis suggest that there is a window of time following primary EBV infection
when there is an increased risk of EBVassociated cHL and that genetic factors, specifically HLA-A genotype, modify this risk.
EBV-associated cHL patients have higher numbers of EBV-infected cells than patients with
EBV-negative disease [112], and infectious
mononucleosis patients have very high numbers
of circulating EBV-­
infected B-cells, which
decrease over time [113]. The number of EBVinfected cells carried by an individual is therefore likely to influence the risk of EBV-associated
cHL. It may, therefore, be possible to decrease
the risk of EBV-positive cHL by EBV vaccination, even in the absence of sterilizing immunity
[114], or by treatment of infectious mononucleosis with antiviral agents.
2.2.3
Epstein-Barr Virus
and Hodgkin Lymphoma:
A Causative Association?
In the absence of good animal models and the
ability to prevent EBV infection, it is difficult to
prove that the association between EBV and
cHL is causal; however, consideration of the
viral, molecular, and epidemiological data pro-
2
The Role of Viruses in the Genesis of Hodgkin Lymphoma
vides support for this idea. (1) The EBV infection in EBV-positive cHL tumors is clonal
indicating that all the tumor cells are derived
from a single EBV-infected cell. (2) In EBVassociated cases, all HRS cells are infected by
the virus. Although EBNA1 facilitates both synchronous replication of the viral episome with
cellular DNA and genome partitioning, this process is not 100% efficient [56]. If the virus were
not required for maintenance of the transformed
phenotype, a gradual loss of viral genomes from
the tumor cells would be anticipated. (3) EBV is
present in the tumor cells of a significant proportion of cHL cases. Although most adults are
infected by EBV, only 1–50 per million B-cells
are EBV-infected in healthy individuals [115].
If EBV were simply a passenger virus, i.e., present in a B-cell that was subsequently transformed by other mechanisms, EBV-associated
cHL would be a rare occurrence. (4) LMP1 and
LMP2A have plausible biological functions in
the pathogenesis of cHL, as described above. (5)
Crippling mutations of immunoglobulin genes
have been described in a quarter of cHL cases,
and almost all of these cases are EBV-­associated
[116]. This is consistent with the idea that EBV
rescues HRS cells (or precursors) that have
destructive mutations of their immunoglobulin
genes from apoptosis. (6) Recent studies show
that EBV-positive cHL has significantly fewer
cellular mutations, including chromosomal
breakpoints and aneuploid autosomes than
EBV-­
negative cHL [117, 118]. Deleterious
mutations of the TNFAIP3 and NFKBIA genes,
which are both negative regulators of NF-κB
signaling, are also much more frequent in HRS
cells from EBV-­negative cases (see Chap. 3)
[119–123]. (7) EBV-­associated cHL cases share
genetic risk factors for disease development,
which are generally distinct from those associated with EBV-negative cHL [101–105, 107,
108, 124, 125]. (8) In some cases, the development of EBV-associated cHL is temporally
related to primary EBV infection [92, 93, 95].
(9) Individuals who subsequently develop EBVassociated cHL have abnormal EBV antibody
profiles before diagnosis [97].
2.2.4
33
Epstein-Barr Virus
and the Clinicopathological
Features of Hodgkin
Lymphoma
Although the above data indicate that EBV-­
positive and EBV-negative cHL have distinct
natural histories, the phenotypic expression of
both processes appears remarkably similar. Gene
expression profiling of HRS cells suggests that
EBV has only a small influence on the transcription profile of established HRS cells [126].
However, EBV status does show clear associations with histological subtype. In a meta-­analysis
of published studies of EBV and cHL, Lee et al.
reported that 66% of MCHL cases are EBV-­
associated, compared to 29% of NSHL cases
[86]. Despite this difference, it is clear that “barn
door” NSHL cases can be EBV-positive, and so
the lack of a complete correlation between histological subtype and EBV status is not simply due
to the criteria used in, and subjective nature of,
histological subtyping. In industrialized countries, NSHL is much more common than MCHL,
and in our experience, the majority (just) of EBV-­
positive cases in the UK are, in fact, NSHL and
not MCHL.
Early studies investigating clinical outcome in
relation to EBV status in cHL appeared conflicting, and the meta-analysis performed by Lee et al.,
which was not able to stratify patients by age, did
not find any associations with survival. However, a
consistent picture has emerged from populationbased studies with age stratification of patients
[127–130]. In young adult patients, there appears
to be no significant difference in overall survival
by EBV status. In contrast, EBV positivity is associated with inferior outcome among patients aged
50 years and over. It is not clear whether this difference is related to the disease process itself or
whether it reflects an underlying comorbidity or
immune dysregulation that potentially predisposes
to EBV-­associated cHL. EBV status is not routinely used in therapeutic decisions, but it is possible that this group of patients would benefit from
alternative treatments, such as third-party cytotoxic lymphocyte infusions or novel therapies tar-
R. F. Jarrett et al.
34
geting EBV. Biomarker levels may also vary by
EBV status; for instance, CCL17 (TARC) levels
are lower in patients with EBV-associated cHL,
but monitoring of plasma EBV levels (a form of
circulating tumor DNA) can be used to assess
treatment response and detect relapse in these
patients [131, 132]. Further studies investigating
these issues are required.
2.3
Epstein-Barr Virus-Negative
Hodgkin Lymphoma Cases
Adolescent and young adult cHL cases are the
group least likely to be associated with EBV,
and yet it is for these cases that there is most
epidemiological evidence suggesting viral
involvement. Early studies reported consistent
associations between young adult HL and correlates of a high standard of living in early
childhood [133]. Many of these associations
with social class variables have not been
detected in recent studies, most probably reflecting societal changes; however, an increased risk
of young adult HL in individuals with less than
1 year of preschool attendance has been
observed [6, 93]. Collectively, the data suggest
that diminished social contact in early childhood is associated with an increased risk of this
disease. Interview and questionnaire data generally support the idea that young adult HL
patients have experienced fewer common infections in childhood [91, 134]. This has led to
speculation that young adult HL is associated
with delayed exposure to one or more common
childhood infections.
A frequent suggestion is that EBV is involved
in all cases of cHL but uses a hit-and-run mechanism in “EBV-negative” cases. This possibility is
very difficult to exclude, but the available data
indicate that it cannot account for all “EBV-­
negative” cases. Importantly, not all cases are
EBV-infected [98, 135]; in fact, we found that
EBV-negative cHL cases in the 15- to 24-year
age group were more likely to be EBV-­
seronegative than age-matched controls [135].
Also, there is no evidence for integration of
incomplete EBV genomes in “EBV-negative”
cHL biopsies [135, 136].
Alternative hypotheses are that lack of exposure to pathogens in early life shapes the microbiome and immune defenses, leading to an
increased risk of developing cHL in young adulthood [137], or that EBV-negative cHL is associated with delayed exposure to another common
virus that is directly involved in disease pathogenesis. Candidate viruses that are common and
have transforming potential include herpesviruses and polyomaviruses. Any virus with a direct
transforming role would be expected to be present in all HRS cells within tumors.
2.3.1
Hodgkin Lymphoma
and Herpesviruses Other Than
Epstein-Barr Virus
At present, there are nine known human herpesviruses (HHVs), including EBV (officially HHV-­
4). All are widespread in distribution, except
herpes simplex virus 2 (HHV-2) and HHV-8.
EBV and Kaposi sarcoma herpesvirus (KSHV,
officially HHV-8) belong to the gammaherpesvirus subfamily of herpesviruses; both infect lymphoid cells and are tumor viruses. KSHV causes
Kaposi sarcoma and rare forms of lymphoma but
is not associated with cHL [138–141]. There is
also no evidence of involvement of the alpha-­
herpesviruses, herpes simplex virus 1, and varicella zoster virus [140]. In contrast, genomes of
the betaherpesviruses, human cytomegalovirus,
HHV-6A, HHV-6B, and HHV-7, have been
detected in cHL tumors using sensitive molecular
assays. Schmidt et al. detected human cytomegalovirus genomes by PCR in 8/86 HL biopsies
[139], although smaller case series failed to identify this virus in tumor samples [140, 142–144].
HHV-7 has been detected in 20–68% of HL biopsies by PCR [139, 140, 144, 145]. However, negative results were obtained using Southern blot
analysis, which is much less sensitive than PCR
but would be expected to detect a virus present in
all HRS cells [146], and there is no evidence that
the virus is present in HRS cells [145]. There is,
2
The Role of Viruses in the Genesis of Hodgkin Lymphoma
therefore, no evidence for direct involvement of
HHV-7 in cHL pathogenesis.
HHV-6 deserves special mention because this
lymphotropic virus has been consistently linked
with cHL. HHV-6 is now classified as two distinct viruses, HHV-6A and HHV-6B [147], rather
than two variants, but until recently many studies
have not distinguished between the two viruses.
Serological studies have shown that HHV-6 antibody titers and, in some studies, seroprevalence
are higher in HL cases than controls [148–150].
We also found that young adults with non-EBV-­
associated HL had higher titers of HHV-6 antibodies
than
age-matched
cases
with
EBV-associated disease (unpublished results).
HHV-7 antibody titers were similar in the two
groups of cases suggesting a specific association
between HHV-6 and cHL.
HHV-6 genomes have been consistently
detected in HL biopsies using PCR although
detection rates range from 8% to 79% [139, 140,
144, 150–155], and some studies have reported
similar detection rates in reactive lymph nodes
[144, 152]. Differences in PCR assay sensitivity
and the amount of DNA assayed most probably
account for the differences in detection rate since
viral genome copy numbers within biopsies are
often low. Up to 87% of NSHL cases have been
reported to be HHV-6-positive [155, 156], but it
is clear that these PCR-positive cases include
both EBV-associated and EBV-non-associated
cases [140, 152, 155, 156]. Both HHV-6A and B
have been detected within biopsies with four
studies showing a clear bias toward HHV-6B
[140, 151, 152, 155], one detecting a higher proportion of HHV-6A-positive tumors [139], and
one detecting HHV-6A and B as well as dual
infections [156]. The low viral genome copy in
many tumors suggests that the virus cannot be
present in every HRS cell and raises the suspicion
that the virus is in cells in the reactive component
of tumors. Very high viral copy numbers must
also be interpreted with caution since inherited
chromosomally integrated HHV-6 (iciHHV-6) is
transmitted in the germline in around 1% of individuals and gives rise to high viral loads since
viral genomes are present in every nucleated cell
35
in the body [157–159]. Following exclusion of
cases with iciHHV-6, studies using the less sensitive technique of Southern blot analysis have
largely been negative suggesting a low viral copy
number within tumors [138, 150, 152, 153, 160].
In contrast, in EBV-associated cHL, EBV
genomes are almost always detectable using this
technique [7, 20, 128]. The critical question is
whether HHV-6 infects HRS cells and, if so, is
the virus present in every HRS cell.
Early studies using in situ hybridization and
IHC reported that the virus was present in cells in
the tumor microenvironment, either exclusively
[152, 161] or with occasional positive HRS cells
[162, 163]. However, two recent studies described
HHV-6-positive HRS cells [156, 164], and we
detected HHV-6 transcripts in an RNAseq analysis of HRS cells enriched from an EBV-negative
cHL biopsy (unpublished data), thus renewing
interest in cHL and HHV-6. Lacroix et al. made a
polyclonal antiserum to the DR7 open reading
frame (ORF) of HHV-6B (designated DR7B) to
examine the cellular localization of the virus in
PCR-positive cases [164]. They selected this particular ORF because the equivalent HHV-6A
ORF has transforming properties and the translated protein binds p53 and inhibits p53-activated
transcription [153, 165]. It is likely that the DR7
ORF is expressed as the second exon of DR6, a
larger nuclear protein [166, 167]. Using this antiserum, cytoplasmic staining of HRS cells was
identified in 28/38 PCR-positive biopsies [164].
In 17 cases, positive staining was exclusive to
HRS cells, and in further 17 cases, positive staining of cells in the microenvironment was noted.
In 15 of the 38 biopsies, HRS cells were also
positively stained using an antibody to the HHV-6
gp116/64/54 glycoprotein. Further analyses suggested that DR7B bound p53, upregulated NF-κB
p105 and p65 promoters, significantly increased
NF-κB activation, and induced upregulation of
Id2. In the second study, Siddon et al. investigated biopsies from 21 NSHL cases, including 18
that were HHV-6-positive by PCR, using multiple approaches [156]. In ten cases, staining of
HRS cells was demonstrated using a commercially available monoclonal antibody raised
R. F. Jarrett et al.
36
against virus lysate (Santa Cruz Biotechnology);
scattered positive HRS cells were also demonstrated using antibodies to the late viral proteins
p41 and p98. Laser capture microdissection
coupled with PCR confirmed the presence of
­
HHV-6 DNA in pooled HRS cells from eight of
the ten IHC-positive biopsies. This study provides the most convincing evidence to date that
HHV-6 can infect HRS cells but does not show
that the virus is present in every HRS cell.
Furthermore, the IHC staining pattern suggests
lytic replication (or abortive replication) rather
than latent infection, and so the outcome of viral
infection in these cells is not clear.
As mentioned above, some individuals
inherit HHV-6 in the germline [157, 159]. The
first study to demonstrate chromosomally integrated HHV-6 investigated three patients with
high viral loads in peripheral blood, including
one with cHL [168, 169]. To determine whether
iciHHV-6 is associated with cHL, we examined
936 cHL cases and 563 controls but found no
evidence that iciHHV-6 was overrepresented
among cases [170].
Overall, the data do not support the idea that
HHV-6 has a direct role in disease pathogenesis.
However, it is possible that HHV-6 is frequently
reactivated in cHL tumors. CD134 is the cellular
receptor for HHV-6B [171], and it is possible that
CD134-positive T-cells in the cHL microenvironment [74] facilitate replication of HHV-6B. Robust
in situ hybridization assays for HHV-6 are
required to confidently rule out a direct role in
cHL.
To search for novel members of the herpesvirus family, we and others have designed degenerate PCR assays which amplify herpesvirus
polymerase and glycoprotein B gene sequences
[140, 172]. The primer sequences in degenerate
assays are derived from well-conserved peptide
motifs in amino acid sequences of proteins;
therefore, these assays should have the ability to
detect genomes from known and currently
unknown viruses [173]. Using herpesvirus polymerase assays, we have not detected novel herpesviruses in cHL biopsies although the assays
had sufficient sensitivity to detect EBV in EBV-­
associated cases, as well as low-level HHV-6
and HHV-7 infection [140] (and unpublished
results).
2.3.2
Polyomaviruses and Hodgkin
Lymphoma
There are now (at least) 14 human polyomaviruses (HPyVs) [174–178]. JCV and BKV were
discovered over 40 years ago, but the others have
all been discovered since 2007 with the advent
of modern molecular techniques for virus discovery. Seroprevalence studies suggest that the
majority of adults are infected by BKPyV,
KIPyV, WUPyV, MCPyV, HPyV6 and 7, and
TSPyV and a significant minority by JCPyV,
HPyV9, and HPyV12 [176, 179–181]. Among
this expanding list of HPyVs, only JCPyV,
BKPyV, TSPyV (associated with trichodysplasia
spinulosa in immunosuppressed persons), and
MCPyV show clear disease associations.
MCPyV is associated with Merkel cell carcinoma and has been categorized by IARC as a
group2A carcinogen (probably carcinogenic to
humans) [175, 182, 183]. It is the only HPyV to
be unambiguously linked with a specific malignancy; however, other polyomaviruses have
oncogenic potential.
Several laboratories have looked for evidence
of HPyV genomes in cHL biopsies. Using sensitive quantitative PCR assays, we found no evidence of JCV or BKV genomes in 35 cHL
biopsies [184]. Hernandez-Losa et al. detected
JCV in 1/20 and BKV in 2/20 cHL samples using
a multiplex, nested PCR [144]. Robles et al.
reported that MCPyV seroprevalence was slightly
higher in HL cases than controls, 84.4% compared to 81.2%, but differences were not statistically significant [185]. Two quantitative PCR
studies detected MCPyV genomes in a small proportion (1/30 and 3/41) of cHL tumors [186,
187]; viral copy numbers were low making it
extremely unlikely that this virus is playing any
role in disease pathogenesis. To date, there have
been no reports on the prevalence of the more
recently identified viruses in cHL.
2
The Role of Viruses in the Genesis of Hodgkin Lymphoma
Degenerate PCR assays have also been applied
to the study of PyVs and HL [184, 187]. Volter
et al. examined five cases of HL using a degenerate PCR assay based on the viral VP1 protein but
did not detect any evidence of polyomavirus infection [187]. We examined 35 cases of cHL, including 23 EBV-negative cases, using 3 ­degenerate
PyV assays based on the large T antigen, and also
obtained negative results [184]. The latter assays
were designed before 2006 and therefore before
most HPyVs were discovered. Alignment of large
T antigen amino acid sequences from the recently
identified viruses suggests that our assays would
be able to detect KIPyV, WUPyV, TSPyV, and
HPyV9 and 10 but not MCPyV, HPyV6, and
HPyV7; however, given the tropism of the latter
viruses for skin, it is unlikely that they are involved
in cHL [176]. Overall, these results provide no evidence for HPyV involvement in the pathogenesis
of cHL, but it remains possible that an unknown
HPyV has escaped detection.
2.3.3
Measles Virus and Hodgkin
Lymphoma
In 2003, Benharroch and colleagues reported an
association between measles virus (MV) and cHL
[188]. They subsequently reported that MV proteins were detectable by IHC in HRS cells from
most HL cases [189]. MV RNA was also detected
by RT-PCR and in situ hybridization in a significant minority of the cases examined [189].
Subsequent studies have failed to confirm these
associations [190, 191]. Our group found no evidence of MV in 97 cHL cases examined by IHC
and 20 cHL cases investigated using RT-PCR
[191]. Similarly, Maggio et al. found no evidence
of MV genomes or transcripts in HRS cells microdissected from biopsies from 18 German and 17
Israeli HL cases [190]; the latter cases had previously scored positive for MV antigens [190].
Epidemiological studies have also failed to show
that MV infection is a risk factor for the development of cHL; on the contrary, the data suggest a
mild protective effect of prior MV infection [91,
134, 192].
2.3.4
37
The Virome, Anelloviruses,
and Hodgkin Lymphoma
It is now recognized that the microbiome, which is
thought to play an important role in shaping the
immune system, includes a large number of viral
species (the virome). Anelloviruses account for
around 70% of these viruses [193]. The anellovirus family includes a large number of genetically
diverse viruses with small, circular, single-­stranded
DNA genomes, which are classified in the Torque
teno virus (TTV), Torque teno midi virus
(TTMDV), and Torque teno mini virus (TTMV)
genera in humans. They are widely distributed,
acquired early in life, and establish persistent
infections, but have not yet been associated with
any disease; however, it has been suggested that
they can modulate both innate and adaptive
immune responses [194]. In 2004, Jelcic et al.
reported the isolation of 24 novel TTVs from a
spleen of an HL patient [195]. This led zur Hausen
and de Villiers to suggest that TTVs could play a
role in the development of leukemias and lymphomas that are associated with a “protected childhood environment” [196]. In their model, they
postulated that TTVs and related anelloviruses
increase the risk of chromosomal abnormalities
and that anellovirus load is increased in individuals who have experienced fewer infections [196].
Increased TTV loads could also contribute to cHL
through modulation of immune defenses. TTVs
have also been identified in cHL tumor biopsies by
other groups [197, 198], but these studies detected
TTVs at a similar frequency in other lymphomas
[197] and reactive nodes [198]. In a recent metagenomic analysis, Pan et al. analyzed the virome in
blood samples from 19 HL patients, 252 nonHodgkin lymphoma patients, and 40 healthy controls from China [199]. Eleven novel, but closely
related, TTMVs were identified in three of the HL
patients but not in the other patients or controls.
The significance of these findings is currently
unclear. Further investigation of the virome in both
cHL patients and individuals with lack of social
contact in early childhood is required to understand the potential contribution of anelloviruses,
the virome, and the microbiome to the risk of cHL.
R. F. Jarrett et al.
38
2.4
Conclusions
5.
While the evidence suggesting a causal relationship between EBV and a proportion of cHL cases
appears strong, current data do not show a consistent and specific association between any virus
and EBV-negative cHL. This does not exclude
viral involvement since the difficulty of obtaining
large numbers of highly enriched HRS cells has
precluded the use of certain techniques, such as
representational difference analysis, in the analysis of cHL [137]. Next-generation sequencing
methods have opened new avenues for virus discovery and have led to the identification of
numerous novel viruses in the last few years
[139, 140, 156]. These techniques provide our
best hope of discovering a new virus in EBV-­
negative HRS cells. It is possible that cellular
mutations substitute for the functions of EBV
genes in EBV-negative HRS cells [126].
Deleterious mutations of inhibitors of the NF-κB
pathway, including genes encoding A20 and
IκBα, appear to be present in the HRS cells of
many cases of EBV-negative cHL (see Chap. 3)
[90–94], and it is possible that these mutations
substitute for LMP1. However, there is no obvious link between these mutations and the epidemiological features of cHL, and involvement of
another virus(es) with either a direct or indirect
role still appears attractive. Understanding the
role of viruses in EBV-negative cHL could potentially open up possibilities for disease prevention
as well as novel therapeutic targets and is a goal
worth pursuing.
17.
References
18.
1. MacMahon B (1957) Epidemiological evidence of the nature of Hodgkin's disease. Cancer
10:1045–1054
2. MacMahon B (1966) Epidemiology of Hodgkin's
disease. Cancer Res 26:1189–1201
3. Gutensohn NM (1982) Social class and age at diagnosis of Hodgkin’s disease: new epidemiologic
evidence for the “two-disease hypothesis”. Cancer
Treat Rep 66:689–695
4. Alexander FE, McKinney PA, Williams J, Ricketts
TJ, Cartwright RA (1991) Epidemiological evidence
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
19.
20.
21.
for the ‘two-disease hypothesis’ in Hodgkin's disease. Int J Epidemiol 20:354–361
Gutensohn NM, Shapiro DS (1982) Social class risk
factors among children with Hodgkin’s disease. Int J
Cancer 30:433–435
Chang ET, Zheng T, Weir EG et al (2004) Childhood
social environment and Hodgkin’s lymphoma: new
findings from a population-based case-control study.
Cancer Epidemiol Biomark Prev 13:1361–1370
Jarrett RF, Gallagher A, Jones DB et al (1991)
Detection of Epstein-Barr virus genomes in
Hodgkin's disease: relation to age. J Clin Pathol
44:844–848
Jarrett RF, Armstrong AA, Alexander E (1996)
Epidemiology of EBV and Hodgkin’s lymphoma.
Ann Oncol 7(Suppl 4):5–10
Glaser SL, Lin RJ, Stewart SL et al (1997) Epstein-­
Barr virus-associated Hodgkin’s disease: epidemiologic characteristics in international data. Int J
Cancer 70:375–382
Longnecker R, Kieff E, Cohen JI (2013) Epstein-­
Barr Virus. In: Fields BN, Knipe DM, Howley PM
(eds) Fields’ virology, 2nd edn. Lippincott Williams
& Wilkins, Philadelphia, PA, pp R1898–RR954
de-The G, Day NE, Geser A et al (1975) Sero-­
epidemiology of the Epstein-Barr virus: preliminary
analysis of an international study – a review. IARC
Sci Publ 11:3–16
Young LS, Yap LF, Murray PG (2016) Epstein-Barr
virus: more than 50 years old and still providing surprises. Nat Rev Cancer 16:789–802
Farrell PJ (2018) Epstein-Barr virus and cancer.
Annu Rev Pathol 14:29–53
Babcock GJ, Decker LL, Volk M, Thorley-Lawson
DA (1998) EBV persistence in memory B cells
in vivo. Immunity 9:395–404
Rickinson AB, Long HM, Palendira U, Munz C,
Hislop AD (2014) Cellular immune controls over
Epstein-Barr virus infection: new lessons from the
clinic and the laboratory. Trends Immunol 35:159–169
Skalsky RL, Cullen BR (2015) EBV noncoding
RNAs. Curr Top Microbiol Immunol 391:181–217
Pfeffer S, Zavolan M, Grasser FA et al (2004)
Identification of virus-encoded microRNAs. Science
304:734–736
Cai X, Schafer A, Lu S et al (2006) Epstein-Barr
virus microRNAs are evolutionarily conserved and
differentially expressed. PLoS Pathog 2:e23
Edwards RH, Marquitz AR, Raab-Traub N (2008)
Epstein-Barr virus BART microRNAs are produced from a large intron prior to splicing. J Virol
82:9094–9106
Zhu JY, Pfuhl T, Motsch N et al (2009) Identification
of novel Epstein-Barr virus microRNA genes from
nasopharyngeal carcinomas. J Virol 83:3333–3341
Cosmopoulos K, Pegtel M, Hawkins J et al (2009)
Comprehensive profiling of Epstein-Barr virus
microRNAs in nasopharyngeal carcinoma. J Virol
83:2357–2367
2
The Role of Viruses in the Genesis of Hodgkin Lymphoma
22. Klinke O, Feederle R, Delecluse HJ (2014) Genetics
of Epstein-Barr virus microRNAs. Semin Cancer
Biol 26:52–59
23. Khanna R, Burrows SR (2000) Role of cytotoxic T
lymphocytes in Epstein-Barr virus-associated diseases. Annu Rev Microbiol 54:19–48
24. Hislop AD, Taylor GS, Sauce D, Rickinson AB
(2007) Cellular responses to viral infection in
humans: lessons from Epstein-Barr virus. Annu Rev
Immunol 25:587–617
25. Pallesen G, Hamilton-Dutoit SJ, Rowe M, Young
LS (1991) Expression of Epstein-Barr virus latent
gene products in tumour cells of Hodgkin’s disease.
Lancet 337:320–322
26. Wu TC, Mann RB, Charache P et al (1990) Detection
of EBV gene expression in Reed-Sternberg cells of
Hodgkin’s disease. Int J Cancer 46:801–804
27. Weiss LM, Movahed LA, Warnke RA, Sklar J (1989)
Detection of Epstein-Barr viral genomes in ReedSternberg cells of Hodgkin’s disease. N Engl J Med
320:502–506
28. Weiss LM, Strickler JG, Warnke RA, Purtilo DT,
Sklar J (1987) Epstein-Barr viral DNA in tissues of
Hodgkin’s disease. Am J Pathol 129:86–91
29. Gledhill S, Gallagher A, Jones DB et al (1991) Viral
involvement in Hodgkin’s disease: detection of
clonal type a Epstein-Barr virus genomes in tumour
samples. Br J Cancer 64:227–232
30. Grasser FA, Murray PG, Kremmer E et al (1994)
Monoclonal antibodies directed against the Epstein-­
Barr virus-encoded nuclear antigen 1 (EBNA1):
immunohistologic detection of EBNA1 in the malignant cells of Hodgkin’s disease. Blood 84:3792–3798
31. Deacon EM, Pallesen G, Niedobitek G et al (1993)
Epstein-Barr virus and Hodgkin’s disease: transcriptional analysis of virus latency in the malignant cells.
J Exp Med 177:339–349
32. Niedobitek G, Kremmer E, Herbst H et al (1997)
Immunohistochemical detection of the Epstein-­
Barr virus-encoded latent membrane protein 2A in
Hodgkin’s disease and infectious mononucleosis.
Blood 90:1664–1672
33. Qiu J, Cosmopoulos K, Pegtel M et al (2011) A
novel persistence associated EBV miRNA expression profile is disrupted in neoplasia. PLoS Pathog
7:e1002193
34. Kuppers R (2009) The biology of Hodgkin’s lymphoma. Nat Rev Cancer 9:15–27
35. Kuppers R (2009) Molecular biology of Hodgkin
lymphoma. Hematology Am Soc Hematol Educ
Program 2009:491–496
36. Kuppers R, Klein U, Schwering I et al (2003)
Identification of Hodgkin and Reed-Sternberg cellspecific genes by gene expression profiling. J Clin
Invest 111:529–537
37. Schwering I, Brauninger A, Klein U et al (2003) Loss
of the B-lineage-specific gene expression program in
Hodgkin and Reed-Sternberg cells of Hodgkin lymphoma. Blood 101:1505–1512
39
38. Bechtel D, Kurth J, Unkel C, Kuppers R (2005)
Transformation of BCR-deficient germinal-center B
cells by EBV supports a major role of the virus in
the pathogenesis of Hodgkin and posttransplantation
lymphomas. Blood 106:4345–4350
39. Mancao C, Altmann M, Jungnickel B,
Hammerschmidt W (2005) Rescue of “crippled”
germinal center B cells from apoptosis by Epstein-­
Barr virus. Blood 106:4339–4344
40. Chaganti S, Bell AI, Pastor NB et al (2005) Epstein-­
Barr virus infection in vitro can rescue germinal center B cells with inactivated immunoglobulin genes.
Blood 106:4249–4252
41. Mancao C, Hammerschmidt W (2007) Epstein-Barr
virus latent membrane protein 2A is a B-cell receptor mimic and essential for B-cell survival. Blood
110:3715–3721
42. Caldwell RG, Brown RC, Longnecker R (2000)
Epstein-Barr virus LMP2A-induced B-cell survival
in two unique classes of EmuLMP2A transgenic
mice. J Virol 74:1101–1113
43. Portis T, Longnecker R (2003) Epstein-Barr virus
LMP2A interferes with global transcription factor
regulation when expressed during B-lymphocyte
development. J Virol 77:105–114
44. Anderson LJ, Longnecker R (2009) Epstein-Barr
virus latent membrane protein 2A exploits Notch1 to
alter B-cell identity in vivo. Blood 113:108–116
45. Portis T, Dyck P, Longnecker R (2003) EpsteinBarr Virus (EBV) LMP2A induces alterations in
gene transcription similar to those observed in
Reed-Sternberg cells of Hodgkin lymphoma. Blood
102:4166–4178
46. Basso K, Klein U, Niu H et al (2004) Tracking CD40
signaling during germinal center development.
Blood 104:4088–4096
47. Devergne O, Cahir McFarland ED, Mosialos G,
Izumi KM, Ware CF, Kieff E (1998) Role of the TRAF
binding site and NF-kappaB activation in Epstein-­
Barr virus latent membrane protein 1-induced cell
gene expression. J Virol 72:7900–7908
48. Izumi KM, Kieff ED (1997) The Epstein-Barr virus
oncogene product latent membrane protein 1 engages
the tumor necrosis factor receptor-associated death
domain protein to mediate B lymphocyte growth
transformation and activate NF-kappaB. Proc Natl
Acad Sci U S A 94:12592–12597
49. Kieser A, Kilger E, Gires O, Ueffing M, Kolch
W, Hammerschmidt W (1997) Epstein-Barr virus
latent membrane protein-1 triggers AP-1 activity
via the c-Jun N-terminal kinase cascade. EMBO J
16:6478–6485
50. Eliopoulos AG, Young LS (1998) Activation of
the cJun N-terminal kinase (JNK) pathway by the
Epstein-Barr virus-encoded latent membrane protein
1 (LMP1). Oncogene 16:1731–1742
51. Eliopoulos AG, Gallagher NJ, Blake SM, Dawson
CW, Young LS (1999) Activation of the p38
mitogen-­
activated protein kinase pathway by
R. F. Jarrett et al.
40
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
Epstein-Barr virus-encoded latent membrane protein 1 c­oregulates interleukin-6 and interleukin-8
production. J Biol Chem 274:16085–16096
Vockerodt M, Morgan SL, Kuo M et al (2008) The
Epstein-Barr virus oncoprotein, latent membrane
protein-1, reprograms germinal centre B cells
towards a Hodgkin’s Reed-Sternberg-like phenotype. J Pathol 216:83–92
Bargou RC, Emmerich F, Krappmann D et al (1997)
Constitutive nuclear factor-kappaB-RelA activation
is required for proliferation and survival of Hodgkin’s
disease tumor cells. J Clin Invest 100:2961–2969
Dutton A, O'Neil JD, Milner AE et al (2004)
Expression of the cellular FLICE-inhibitory protein
(c-FLIP) protects Hodgkin’s lymphoma cells from
autonomous Fas-mediated death. Proc Natl Acad Sci
U S A 101:6611–6616
Kashkar H, Haefs C, Shin H et al (2003) XIAP-­
mediated caspase inhibition in Hodgkin’s
lymphoma-­derived B cells. J Exp Med 198:341–347
Nanbo A, Sugden A, Sugden B (2007) The coupling of synthesis and partitioning of EBV’s plasmid replicon is revealed in live cells. EMBO J
26:4252–4262
Kennedy G, Komano J, Sugden B (2003) Epstein-­
Barr virus provides a survival factor to Burkitt’s lymphomas. Proc Natl Acad Sci U S A 100:14269–14274
Wilson JB, Bell JL, Levine AJ (1996) Expression of
Epstein-Barr virus nuclear antigen-1 induces B cell
neoplasia in transgenic mice. EMBO J 15:3117–3126
Kang MS, Lu H, Yasui T et al (2005) Epstein-Barr
virus nuclear antigen 1 does not induce lymphoma
in transgenic FVB mice. Proc Natl Acad Sci U S A
102:820–825
Kang MS, Soni V, Bronson R, Kieff E (2008)
Epstein-Barr virus nuclear antigen 1 does not cause
lymphoma in C57BL/6J mice. J Virol 82:4180–4183
Yajima M, Kanda T, Takada K (2005) Critical role of
Epstein-Barr Virus (EBV)-encoded RNA in efficient
EBV-induced B-lymphocyte growth transformation.
J Virol 79:4298–4307
Skalsky RL, Corcoran DL, Gottwein E et al (2012)
The viral and cellular microRNA targetome in lymphoblastoid cell lines. PLoS Pathog 8:e1002484
Hancock MH, Skalsky RL (2018) Roles of non-­
coding RNAs during herpesvirus infection. Curr Top
Microbiol Immunol 419:243–280
Albanese M, Tagawa T, Buschle A, Hammerschmidt
W (2017) MicroRNAs of Epstein-Barr virus control innate and adaptive antiviral immunity. J Virol
91:pii: e01667
Chen Y, Fachko D, Ivanov NS, Skinner CM, Skalsky
RL (2019) Epstein-Barr virus microRNAs regulate
B cell receptor signal transduction and lytic reactivation. PLoS Pathog 15:e1007535
Murer A, Ruhl J, Zbinden A et al (2019) MicroRNAs
of Epstein-Barr virus attenuate T-cell-mediated
immune control in vivo. MBio 10:e01941–e01918
Ross N, Gandhi MK, Nourse JP (2013) The Epstein-­
Barr virus microRNA BART11-5p targets the early
68.
69.
70.
71.
72.
73.
74.
75.
76.
77.
78.
79.
80.
B-cell transcription factor EBF1. Am J Blood Res
3:210–224
Godshalk SE, Bhaduri-McIntosh S, Slack FJ (2008)
Epstein-Barr virus-mediated dysregulation of human
microRNA expression. Cell Cycle 7:3595–3600
van den Berg A, Kroesen BJ, Kooistra K et al (2003)
High expression of B-cell receptor inducible gene
BIC in all subtypes of Hodgkin lymphoma. Genes
Chromosomes Cancer 37:20–28
Navarro A, Gaya A, Martinez A et al (2008)
MicroRNA expression profiling in classic Hodgkin
lymphoma. Blood 111:2825–2832
Vrazo AC, Chauchard M, Raab-Traub N,
Longnecker R (2012) Epstein-Barr virus LMP2A
reduces hyperactivation induced by LMP1 to restore
normal B cell phenotype in transgenic mice. PLoS
Pathog 8:e1002662
Vrzalikova K, Ibrahim M, Nagy E et al (2018)
Co-expression of the Epstein-Barr Virus-encoded
latent membrane proteins and the pathogenesis of
classic Hodgkin lymphoma. Cancers (Basel) 10:285
Wirtz T, Weber T, Kracker S, Sommermann T,
Rajewsky K, Yasuda T (2016) Mouse model for
acute Epstein-Barr virus infection. Proc Natl Acad
Sci U S A 113:13821–13826
Greaves P, Clear A, Owen A et al (2013) Defining
characteristics of classical Hodgkin lymphoma microenvironment T-helper cells. Blood
122:2856–2863
Morales O, Mrizak D, Francois V et al (2014)
Epstein-Barr virus infection induces an increase
of T regulatory type 1 cells in Hodgkin lymphoma
patients. Br J Haematol 166:875–890
Oudejans JJ, Jiwa NM, Kummer JA et al (1996)
Analysis of major histocompatibility complex class
I expression on Reed-Sternberg cells in relation to
the cytotoxic T-cell response in Epstein-Barr viruspositive and -negative Hodgkin’s disease. Blood
87:3844–3851
Barros MH, Segges P, Vera-Lozada G, Hassan R,
Niedobitek G (2015) Macrophage polarization
reflects T cell composition of tumor microenvironment in pediatric classical Hodgkin lymphoma and
has impact on survival. PLoS One 10:e0124531
Kis LL, Takahara M, Nagy N, Klein G, Klein E
(2006) Cytokine mediated induction of the major
Epstein-Barr virus (EBV)-encoded transforming
protein, LMP-1. Immunol Lett 104:83–88
Dukers DF, Jaspars LH, Vos W et al (2000)
Quantitative immunohistochemical analysis of
cytokine profiles in Epstein-Barr virus-positive
and -negative cases of Hodgkin’s disease. J Pathol
190:143–149
Khanna R, Burrows SR, Nicholls J, Poulsen LM
(1998) Identification of cytotoxic T cell epitopes
within Epstein-Barr virus (EBV) oncogene latent
membrane protein 1 (LMP1): evidence for HLA A2
supertype-restricted immune recognition of EBV-­
infected cells by LMP1-specific cytotoxic T lymphocytes. Eur J Immunol 28:451–458
2
The Role of Viruses in the Genesis of Hodgkin Lymphoma
81. Lee SP, Thomas WA, Murray RJ et al (1993) HLA
A2.1-restricted cytotoxic T cells recognizing a range
of Epstein-Barr virus isolates through a defined
epitope in latent membrane protein LMP2. J Virol
67:7428–7435
82. Green MR, Rodig S, Juszczynski P et al (2012)
Constitutive AP-1 activity and EBV infection induce
PD-L1 in Hodgkin lymphomas and posttransplant
lymphoproliferative disorders: implications for targeted therapy. Clin Cancer Res 18:1611–1618
83. Nakagomi H, Dolcetti R, Bejarano MT, Pisa P,
Kiessling R, Masucci MG (1994) The Epstein-Barr
virus latent membrane protein-1 (LMP1) induces
interleukin-10 production in Burkitt lymphoma
lines. Int J Cancer 57:240–244
84. Cader FZ, Vockerodt M, Bose S et al (2013) The
EBV oncogene LMP1 protects lymphoma cells from
cell death through the collagen-mediated activation
of DDR1. Blood 122:4237–4245
85. Jarrett RF, Krajewski AS, Angus B et al (2003) The
Scotland and Newcastle epidemiological study of
Hodgkin’s disease: impact of histopathological
review and EBV status on incidence estimates. J
Clin Pathol 56:811–816
86. Lee JH, Kim Y, Choi JW, Kim YS (2014) Prevalence
and prognostic significance of Epstein-Barr virus
infection in classical Hodgkin’s lymphoma: a meta-­
analysis. Arch Med Res 45:417–431
87. Armstrong AA, Alexander FE, Paes RP et al (1993)
Association of Epstein-Barr virus with pediatric
Hodgkin’s disease. Am J Pathol 142:1683–1688
88. Flavell K, Constandinou C, Lowe D et al (1999)
Effect of material deprivation on Epstein-Barr virus
infection in Hodgkin’s disease in the west midlands.
Br J Cancer 80:604–608
89. Henle G, Henle W, Clifford P et al (1969) Antibodies
to Epstein-Barr virus in Burkitt’s lymphoma and
control groups. J Natl Cancer Inst 43:1147–1157
90. Crawford DH, Macsween KF, Higgins CD et al
(2006) A cohort study among university students:
identification of risk factors for Epstein-Barr virus
seroconversion and infectious mononucleosis. Clin
Infect Dis 43:276–282
91. Alexander FE, Jarrett RF, Lawrence D et al (2000)
Risk factors for Hodgkin’s disease by Epstein-Barr
virus (EBV) status: prior infection by EBV and other
agents. Br J Cancer 82:1117–1121
92. Hjalgrim H, Askling J, Rostgaard K et al (2003)
Characteristics of Hodgkin’s lymphoma after
infectious mononucleosis. N Engl J Med
349:1324–1332
93. Hjalgrim H, Smedby KE, Rostgaard K et al (2007)
Infectious mononucleosis, childhood social environment, and risk of Hodgkin lymphoma. Cancer Res
67:2382–2388
94. Glaser SL, Clarke CA, Gulley ML et al (2003)
Population-based patterns of human immunodeficiency virus-related Hodgkin lymphoma in the
greater San Francisco Bay Area, 1988–1998. Cancer
98:300–309
41
95. Quinlan SC, Landgren O, Morton LM, Engels
EA (2010) Hodgkin lymphoma among US solid
organ transplant recipients. Transplantation
90:1011–1015
96. Jarrett RF (2002) Viruses and Hodgkin’s lymphoma.
Ann Oncol 13(Suppl 1):23–29
97. Levin LI, Chang ET, Ambinder RF et al (2012)
Atypical prediagnosis Epstein-Barr virus serology
restricted to EBV-positive Hodgkin lymphoma.
Blood 120:3750–3755
98. Chang ET, Zheng T, Lennette ET et al (2004)
Heterogeneity of risk factors and antibody profiles
in Epstein-Barr virus genome-positive and -negative
Hodgkin lymphoma. J Infect Dis 189:2271–2281
99. Henle W, Henle G, Andersson J et al (1987) Antibody
responses to Epstein-Barr virus-determined nuclear
antigen (EBNA)-1 and EBNA-2 in acute and chronic
Epstein-Barr virus infection. Proc Natl Acad Sci U S
A 84:570–574
100. Rubicz R, Yolken R, Drigalenko E et al (2013)
A genome-wide integrative genomic study localizes genetic factors influencing antibodies against
Epstein-Barr virus nuclear antigen 1 (EBNA-1).
PLoS Genet 9:e1003147
101. Diepstra A, Niens M, Vellenga E et al (2005)
Association with HLA class I in Epstein-Barr-­
virus-positive and with HLA class III in Epstein-­
Barr-­
virus-negative Hodgkin’s lymphoma. Lancet
365:2216–2224
102. Niens M, Jarrett RF, Hepkema B et al (2007) HLA-­
A∗02 is associated with a reduced risk and HLA-­
A∗01 with an increased risk of developing EBV+
Hodgkin lymphoma. Blood 110:3310–3315
103. Hjalgrim H, Rostgaard K, Johnson PC et al (2010)
HLA-A alleles and infectious mononucleosis suggest a critical role for cytotoxic T-cell response in
EBV-related Hodgkin lymphoma. Proc Natl Acad
Sci U S A 107:6400–6405
104. Urayama KY, Jarrett RF, Hjalgrim H et al (2012)
Genome-wide association study of classical
Hodgkin lymphoma and Epstein-Barr virus status-­
defined subgroups. J Natl Cancer Inst 104:240–253
105. Huang X, Kushekhar K, Nolte I et al (2012) HLA
associations in classical Hodgkin lymphoma: EBV
status matters. PLoS One 7:e39986
106. Huang X, Hepkema B, Nolte I et al (2012) HLA-­
A∗02:07 is a protective allele for EBV negative and
a susceptibility allele for EBV positive classical
Hodgkin lymphoma in China. PLoS One 7:e31865
107. Johnson PC, McAulay KA, Montgomery D et al
(2015) Modeling HLA associations with EBV-­
positive and -negative Hodgkin lymphoma suggests
distinct mechanisms in disease pathogenesis. Int J
Cancer 137:1066–1075
108. Delahaye-Sourdeix M, Urayama KY, Gaborieau V
et al (2015) A novel risk locus at 6p21.3 for Epstein-­
Barr virus-positive Hodgkin lymphoma. Cancer
Epidemiol Biomark Prev 24:1838–1843
109. Brennan RM, Burrows SR (2008) A mechanism for
the HLA-A∗01-associated risk for EBV+ Hodgkin
R. F. Jarrett et al.
42
110.
111.
112.
113.
114.
115.
116.
117.
118.
119.
120.
121.
122.
123.
lymphoma and infectious mononucleosis. Blood
112:2589–2590
Alexander FE, Lawrence DJ, Freeland J et al (2003)
An epidemiologic study of index and f­amily infectious mononucleosis and adult Hodgkin’s disease (HD): evidence for a specific association
with EBV+ve HD in young adults. Int J Cancer
107:298–302
McAulay KA, Higgins CD, Macsween KF et al
(2007) HLA class I polymorphisms are associated
with development of infectious mononucleosis upon
primary EBV infection. J Clin Invest 117:3042–3048
Khan G, Lake A, Shield L et al (2005) Phenotype
and frequency of Epstein-Barr virus-infected cells
in pretreatment blood samples from patients with
Hodgkin lymphoma. Br J Haematol 129:511–519
Hochberg D, Souza T, Catalina M, Sullivan JL,
Luzuriaga K, Thorley-Lawson DA (2004) Acute
infection with Epstein-Barr virus targets and overwhelms the peripheral memory B-cell compartment with resting, latently infected cells. J Virol
78:5194–5204
Cohen JI, Mocarski ES, Raab-Traub N, Corey L,
Nabel GJ (2013) The need and challenges for development of an Epstein-Barr virus vaccine. Vaccine
31(Suppl 2):B194–B196
Khan G, Miyashita EM, Yang B, Babcock GJ,
Thorley-Lawson DA (1996) Is EBV persistence
in vivo a model for B cell homeostasis? Immunity
5:173–179
Brauninger A, Schmitz R, Bechtel D, Renne C,
Hansmann ML, Kuppers R (2006) Molecular biology of Hodgkin’s and Reed/Sternberg cells in
Hodgkin’s lymphoma. Int J Cancer 118:1853–1861
Montgomery ND, Coward WB, Johnson S et al
(2016) Karyotypic abnormalities associated with
Epstein-Barr virus status in classical Hodgkin lymphoma. Cancer Genet 209:408–416
Tiacci E, Ladewig E, Schiavoni G et al (2018)
Pervasive mutations of JAK-STAT pathway genes in
classical Hodgkin lymphoma. Blood 131:2454–2465
Schmitz R, Hansmann ML, Bohle V et al (2009)
TNFAIP3 (A20) is a tumor suppressor gene in
Hodgkin lymphoma and primary mediastinal B cell
lymphoma. J Exp Med 206:981–989
Cabannes E, Khan G, Aillet F, Jarrett RF, Hay RT
(1999) Mutations in the IkBa gene in Hodgkin’s
disease suggest a tumour suppressor role for
IkappaBalpha. Oncogene 18:3063–3070
Emmerich F, Meiser M, Hummel M et al (1999)
Overexpression of I kappa B alpha without inhibition of NF-kappaB activity and mutations in the I
kappa B alpha gene in Reed-Sternberg cells. Blood
94:3129–3134
Jungnickel B, Staratschek-Jox A, Brauninger
A et al (2000) Clonal deleterious mutations in
the IkappaBalpha gene in the malignant cells in
Hodgkin’s lymphoma. J Exp Med 191:395–402
Lake A, Shield LA, Cordano P et al (2009) Mutations
of NFKBIA, encoding IkappaB alpha, are a recur-
124.
125.
126.
127.
128.
129.
130.
131.
132.
133.
134.
135.
136.
rent finding in classical Hodgkin lymphoma but are
not a unifying feature of non-EBV-associated cases.
Int J Cancer 125:1334–1342
Enciso-Mora V, Broderick P, Ma Y et al (2010) A
genome-wide association study of Hodgkin’s lymphoma identifies new susceptibility loci at 2p16.1
(REL), 8q24.21 and 10p14 (GATA3). Nat Genet
42:1126–1130
Cozen W, Timofeeva MN, Li D et al (2014) A meta-­
analysis of Hodgkin lymphoma reveals 19p13.3
TCF3 as a novel susceptibility locus. Nat Commun
5:3856
Tiacci E, Doring C, Brune V et al (2012) Analyzing
primary Hodgkin and Reed-Sternberg cells to capture the molecular and cellular pathogenesis of classical Hodgkin lymphoma. Blood 120:4609–4620
Clarke CA, Glaser SL, Dorfman RF et al (2001)
Epstein-Barr virus and survival after Hodgkin disease in a population-based series of women. Cancer
91:1579–1587
Jarrett RF, Stark GL, White J et al (2005) Impact
of tumor Epstein-Barr virus status on presenting
features and outcome in age-defined subgroups
of patients with classic Hodgkin lymphoma: a
population-­based study. Blood 106:2444–2451
Keegan TH, Glaser SL, Clarke CA et al (2005)
Epstein-Barr virus as a marker of survival after
Hodgkin’s lymphoma: a population-based study. J
Clin Oncol 23:7604–7613
Diepstra A, van Imhoff GW, Schaapveld M et al
(2009) Latent Epstein-Barr virus infection of tumor
cells in classical Hodgkin’s lymphoma predicts
adverse outcome in older adult patients. J Clin Oncol
27:3815–3821
Gallagher A, Armstrong AA, MacKenzie J et al
(1999) Detection of Epstein-Barr virus (EBV)
genomes in the serum of patients with EBV-­
associated Hodgkin’s disease. Int J Cancer
84:442–448
Kanakry J, Ambinder R (2015) The biology and
clinical utility of EBV monitoring in blood. Curr Top
Microbiol Immunol 391:475–499
Gutensohn N, Cole P (1977) Epidemiology of
Hodgkin’s disease in the young. Int J Cancer
19:595–604
Glaser SL, Keegan TH, Clarke CA et al (2005)
Exposure to childhood infections and risk of
Epstein-Barr virus--defined Hodgkin’s lymphoma in
women. Int J Cancer 115:599–605
Gallagher A, Perry J, Freeland J et al (2003)
Hodgkin lymphoma and Epstein-Barr virus (EBV):
no evidence to support hit-and-run mechanism in
cases classified as non-EBV-associated. Int J Cancer
104:624–630
Staratschek-Jox A, Kotkowski S, Belge G et al
(2000) Detection of Epstein-Barr virus in HodgkinReed-Sternberg cells: no evidence for the persistence
of integrated viral fragments in latent membrane
protein-1 (LMP-1)-negative classical Hodgkin’s disease. Am J Pathol 156:209–216
2
The Role of Viruses in the Genesis of Hodgkin Lymphoma
137. Cozen W, Yu G, Gail MH et al (2013) Fecal microbiota diversity in survivors of adolescent/young adult
Hodgkin lymphoma: a study of twins. Br J Cancer
108:1163–1167
138. Armstrong AA, Shield L, Gallagher A, Jarrett
RF (1998) Lack of involvement of known oncogenic DNA viruses in Epstein-Barr virus-negative
Hodgkin’s disease. Br J Cancer 77:1045–1047
139. Schmidt CA, Oettle H, Peng R et al (2000) Presence
of human beta- and gamma-herpes virus DNA in
Hodgkin’s disease. Leuk Res 24:865–870
140. Gallagher A, Perry J, Shield L, Freeland J,
MacKenzie J, Jarrett RF (2002) Viruses and
Hodgkin disease: no evidence of novel herpesviruses in non-­EBV-­associated lesions. Int J Cancer
101:259–264
141. Benavente Y, Mbisa G, Labo N et al (2011)
Antibodies against lytic and latent Kaposi’s sarcoma-­
associated herpes virus antigens and lymphoma in
the European EpiLymph case-control study. Br J
Cancer 105:1768–1771
142. Samoszuk M, Ravel J (1991) Frequent detection
of Epstein-Barr viral deoxyribonucleic acid and
absence of cytomegalovirus deoxyribonucleic acid
in Hodgkin’s disease and acquired immunodeficiency syndrome-related Hodgkin’s disease. Lab
Investig 65:631–636
143. Lin SH, Yeh HM, Tzeng CH, Chen PM (1993)
Immunoglobulin and T cell receptor beta chain gene
rearrangements and Epstein-Barr viral DNA in tissues of Hodgkin’s disease in Taiwan. Int J Hematol
57:251–257
144. Hernandez-Losa J, Fedele CG, Pozo F et al (2005)
Lack of association of polyomavirus and herpesvirus types 6 and 7 in human lymphomas. Cancer
103:293–298
145. Secchiero P, Bonino LD, Lusso P et al (1998)
Human herpesvirus type 7 in Hodgkin’s disease. Br
J Haematol 101:492–499
146. Berneman ZN, Torelli G, Luppi M, Jarrett RF (1998)
Absence of a directly causative role for human herpesvirus 7 in human lymphoma and a review of
human herpesvirus 6 in human malignancy. Ann
Hematol 77:275–278
147. Ablashi D, Agut H, Alvarez-Lafuente R et al (2014)
Classification of HHV-6A and HHV-6B as distinct
viruses. Arch Virol 159:863–870
148. Ablashi DV, Josephs SF, Buchbinder A et al (1988)
Human B-lymphotropic virus (human herpesvirus­6). J Virol Methods 21:29–48
149. Clark DA, Alexander FE, McKinney PA et al (1990)
The seroepidemiology of human herpesvirus-6
(HHV-6) from a case-control study of leukaemia and
lymphoma. Int J Cancer 45:829–833
150. Torelli G, Marasca R, Luppi M et al (1991) Human
herpesvirus-6 in human lymphomas: identification
of specific sequences in Hodgkin’s lymphomas by
polymerase chain reaction. Blood 77:2251–2258
151. Di Luca D, Dolcetti R, Mirandola P et al (1994)
Human herpesvirus 6: a survey of presence and
152.
153.
154.
155.
156.
157.
158.
159.
160.
161.
162.
163.
164.
165.
43
variant distribution in normal peripheral lymphocytes and lymphoproliferative disorders. J Infect Dis
170:211–215
Valente G, Secchiero P, Lusso P et al (1996) Human
herpesvirus 6 and Epstein-Barr virus in Hodgkin’s
disease: a controlled study by polymerase chain
reaction and in situ hybridization. Am J Pathol
149:1501–1510
Kashanchi F, Araujo J, Doniger J et al (1997) Human
herpesvirus 6 (HHV-6) ORF-1 transactivating gene
exhibits malignant transforming activity and its protein binds to p53. Oncogene 14:359–367
Collot S, Petit B, Bordessoule D et al (2002) Real-­
time PCR for quantification of human herpesvirus 6
DNA from lymph nodes and saliva. J Clin Microbiol
40:2445–2451
Lacroix A, Jaccard A, Rouzioux C et al (2007)
HHV-6 and EBV DNA quantitation in lymph nodes
of 86 patients with Hodgkin’s lymphoma. J Med
Virol 79:1349–1356
Siddon A, Lozovatsky L, Mohamed A, Hudnall
SD (2012) Human herpesvirus 6 positive ReedSternberg cells in nodular sclerosis Hodgkin lymphoma. Br J Haematol 158:635–643
Daibata M, Taguchi T, Nemoto Y, Taguchi H, Miyoshi
I (1999) Inheritance of chromosomally integrated
human herpesvirus 6 DNA. Blood 94:1545–1549
Leong HN, Tuke PW, Tedder RS et al (2007) The
prevalence of chromosomally integrated human
herpesvirus 6 genomes in the blood of UK blood
donors. J Med Virol 79:45–51
Kaufer BB, Flamand L (2014) Chromosomally integrated HHV-6: impact on virus, cell and organismal
biology. Curr Opin Virol 9:111–118
Luppi M, Barozzi P, Marasca R, Ceccherini-Nelli
L, Torelli G (1993) Characterization of human herpesvirus 6 genomes from cases of latent infection in
human lymphomas and immune disorders. J Infect
Dis 168:1074–1075
Maeda A, Sata T, Enzan H et al (1993) The evidence
of human herpesvirus 6 infection in the lymph nodes
of Hodgkin’s disease. Virchows Arch A Pathol Anat
Histopathol 423:71–75
Rojo J, Ferrer Argote VE, Klueppelberg U et al
(1994) Semi-quantitative in situ hybridization and
immunohistology for antigen expression of human
herpesvirus-6 in various lymphoproliferative diseases. In Vivo 8:517–526
Luppi M, Barozzi P, Garber R et al (1998) Expression
of human herpesvirus-6 antigens in benign and
malignant lymphoproliferative diseases. Am J Pathol
153:815–823
Lacroix A, Collot-Teixeira S, Mardivirin L et al
(2010) Involvement of human herpesvirus-6 variant
B in classic Hodgkin’s lymphoma via DR7 oncoprotein. Clin Cancer Res 16:4711–4721
Thompson J, Choudhury S, Kashanchi F et al (1994)
A transforming fragment within the direct repeat
region of human herpesvirus type 6 that transactivates HIV-1. Oncogene 9:1167–1175
44
166. Schleimann MH, Hoberg S, Solhoj Hansen A et al
(2014) The DR6 protein from human herpesvirus6B induces p53-independent cell cycle arrest in
­
G2/M. Virology 452-453:254–263
167. Megaw AG, Rapaport D, Avidor B, Frenkel N,
Davison AJ (1998) The DNA sequence of the RK
strain of human herpesvirus 7. Virology 244:119–132
168. Luppi M, Marasca R, Barozzi P et al (1993) Three
cases of human herpesvirus-6 latent infection: integration of viral genome in peripheral blood mononuclear cell DNA. J Med Virol 40:44–52
169. Torelli G, Barozzi P, Marasca R et al (1995) Targeted
integration of human herpesvirus 6 in the p arm of
chromosome 17 of human peripheral blood mononuclear cells in vivo. J Med Virol 46:178–188
170. Bell AJ, Gallagher A, Mottram T et al (2014) Germ-­
line transmitted, chromosomally integrated HHV-6
and classical Hodgkin lymphoma. PLoS One
9:e112642
171. Tang H, Serada S, Kawabata A et al (2013) CD134
is a cellular receptor specific for human herpesvirus­6B entry. Proc Natl Acad Sci U S A 110:9096–9099
172. Ehlers B, Borchers K, Grund C, Frolich K, Ludwig H,
Buhk HJ (1999) Detection of new DNA polymerase
genes of known and potentially novel herpesviruses by PCR with degenerate and deoxyinosine-­
substituted primers. Virus Genes 18:211–220
173. Jarrett RF, Johnson D, Wilson KS, Gallagher A
(2006) Molecular methods for virus discovery. Dev
Biol (Basel) 123:77–88. discussion 119–132
174. Allander T, Andreasson K, Gupta S et al (2007)
Identification of a third human polyomavirus. J Virol
81:4130–4136
175. Feng H, Shuda M, Chang Y, Moore PS (2008) Clonal
integration of a polyomavirus in human Merkel cell
carcinoma. Science 319:1096–1100
176. Ehlers B, Wieland U (2013) The novel human
polyomaviruses HPyV6, 7, 9 and beyond. APMIS
121:783–795
177. Gaynor AM, Nissen MD, Whiley DM et al (2007)
Identification of a novel polyomavirus from patients
with acute respiratory tract infections. PLoS Pathog
3:e64
178. Prado JCM, Monezi TA, Amorim AT, Lino V,
Paladino A, Boccardo E (2018) Human polyomaviruses and cancer: an overview. Clinics (Sao Paulo)
73:e558s
179. Knowles WA, Pipkin P, Andrews N et al (2003)
Population-based study of antibody to the human
polyomaviruses BKV and JCV and the simian polyomavirus SV40. J Med Virol 71:115–123
180. Kean JM, Rao S, Wang M, Garcea RL (2009)
Seroepidemiology of human polyomaviruses. PLoS
Pathog 5:e1000363
181. Tolstov YL, Pastrana DV, Feng H et al (2009)
Human Merkel cell polyomavirus infection II. MCV
is a common human infection that can be detected by
conformational capsid epitope immunoassays. Int J
Cancer 125:1250–1256
R. F. Jarrett et al.
182. Kassem A, Schopflin A, Diaz C et al (2008) Frequent
detection of Merkel cell polyomavirus in human
Merkel cell carcinomas and identification of a unique
deletion in the VP1 gene. Cancer Res 68:5009–5013
183. IARC (2014) Malaria and some polyomaviruses
(SV40, BK, JC, and Merkel cell viruses). IARC
Monogr Eval Carcinog Risks Hum 104:9–350
184. Wilson KS, Gallagher A, Freeland JM, Shield LA,
Jarrett RF (2006) Viruses and Hodgkin lymphoma:
no evidence of polyomavirus genomes in tumor
biopsies. Leuk Lymphoma 47:1315–1321
185. Robles C, Poloczek A, Casabonne D et al (2012)
Antibody response to Merkel cell polyomavirus
associated with incident lymphoma in the Epilymph
case-control study in Spain. Cancer Epidemiol
Biomark Prev 21:1592–1598
186. Shuda M, Arora R, Kwun HJ et al (2009) Human
Merkel cell polyomavirus infection I. MCV T
antigen expression in Merkel cell carcinoma, lymphoid tissues and lymphoid tumors. Int J Cancer
125:1243–1249
187. Volter C, Hausen H, Alber D, de Villiers EM (1997)
Screening human tumor samples with a broad-­
spectrum polymerase chain reaction method for the
detection of polyomaviruses. Virology 237:389–396
188. Benharroch D, Shemer-Avni Y, Levy A et al (2003)
New candidate virus in association with Hodgkin’s
disease. Leuk Lymphoma 44:605–610
189. Benharroch D, Shemer-Avni Y, Myint YY et al
(2004) Measles virus: evidence of an association
with Hodgkin’s disease. Br J Cancer 91:572–579
190. Maggio E, Benharroch D, Gopas J, Dittmer U,
Hansmann ML, Kuppers R (2007) Absence of measles virus genome and transcripts in Hodgkin-Reed/
Sternberg cells of a cohort of Hodgkin lymphoma
patients. Int J Cancer 121:448–453
191. Wilson KS, Freeland JM, Gallagher A et al (2007)
Measles virus and classical Hodgkin lymphoma:
no evidence for a direct association. Int J Cancer
121:442–447
192. Karunanayake CP, Singh GV, Spinelli JJ et al (2009)
Occupational exposures and Hodgkin lymphoma:
Canadian case-control study. J Occup Environ Med
51:1447–1454
193. De Vlaminck I, Khush KK, Strehl C et al (2013)
Temporal response of the human virome to
immunosuppression and antiviral therapy. Cell
155:1178–1187
194. Freer G, Maggi F, Pifferi M, Di Cicco ME, Peroni
DG, Pistello M (2018) The virome and its major
component, Anellovirus, a convoluted system molding human immune defenses and possibly affecting
the development of asthma and respiratory diseases
in childhood. Front Microbiol 9:686
195. Jelcic I, Hotz-Wagenblatt A, Hunziker A, Zur
Hausen H, de Villiers EM (2004) Isolation of multiple TT virus genotypes from spleen biopsy tissue
from a Hodgkin’s disease patient: genome reorganization and diversity in the hypervariable region. J
Virol 78:7498–7507
2
The Role of Viruses in the Genesis of Hodgkin Lymphoma
196. zur Hausen H, de Villiers EM (2005) Virus target cell
conditioning model to explain some epidemiologic
characteristics of childhood leukemias and lymphomas. Int J Cancer 115:1–5
197. Garbuglia AR, Iezzi T, Capobianchi MR et al (2003)
Detection of TT virus in lymph node biopsies of
B-cell lymphoma and Hodgkin’s disease, and its
association with EBV infection. Int J Immunopathol
Pharmacol 16:109–118
45
198. Figueiredo CP, Franz-Vasconcelos HC, Giunta G et
al (2007) Detection of Torque Teno virus in EpsteinBarr virus positive and negative lymph nodes of
patients with Hodgkin lymphoma. Leuk Lymphoma
48:731–735
199. Pan S, Yu T, Wang Y et al (2018) Identification of a
Torque Teno Mini Virus (TTMV) in Hodgkin’s lymphoma patients. Front Microbiol 9:1680
3
Pathology and Molecular
Pathology of Hodgkin Lymphoma
Andreas Rosenwald and Ralf Küppers
Contents
3.1 Subclassification and Pathology 3.1.1 Nodular Lymphocyte-­Predominant Hodgkin Lymphoma 3.1.2 Classical Hodgkin Lymphoma: The HRS Cells 3.1.2.1 Nodular Sclerosis Classical Hodgkin Lymphoma 3.1.2.2 Mixed Cellularity Classical Hodgkin Lymphoma 3.1.2.3 Lymphocyte-Depleted Classical Hodgkin Lymphoma 3.1.2.4 Lymphocyte-Rich Classical Hodgkin Lymphoma 47
48
50
51
51
51
51
3.2 Differential Diagnosis 52
3.3 Histogenesis of HRS and LP Cells 3.3.1 Cellular Origin of HRS and LP Cells 3.3.2 Relationship of Hodgkin Cells and Reed-Sternberg Cells and Putative
HRS Cell Precursors 52
52
3.4 Genetic Lesions 54
3.5 Deregulated Transcription Factor Networks and Signaling
Pathways 3.5.1 The Lost B Cell Phenotype 3.5.2 Constitutive Activation of Multiple Signaling Pathways 57
57
59
3.6 Anti-apoptotic Mechanisms 60
References 61
3.1
A. Rosenwald
Institute of Pathology, University of Würzburg,
Würzburg, Germany
e-mail: rosenwald@mail.uni-wuerzburg.de
R. Küppers (*)
Institute of Cell Biology (Cancer Research), Medical
School, University of Duisburg-Essen,
Essen, Germany
e-mail: ralf.kueppers@uk-essen.de
54
Subclassification
and Pathology
The history of Hodgkin lymphoma (HL) dates
back to the first half of the nineteenth century (see
Chap. 1), and it has also been an established view
for quite some time that HL comprises two different disease entities, namely, classical Hodgkin
lymphoma (cHL) and nodular lymphocyte-­
predominant Hodgkin lymphoma (LPHL) [1].
Both entities have in common that the neoplastic
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_3
47
A. Rosenwald and R. Küppers
48
cell population, which can be mononucleated or
multinucleated, makes up only a small percentage of all cells present in an affected lymph node.
However, morphological, clinical, epidemiologic,
and molecular evidence strongly support the
belief that the pathogenesis of these lymphomas
is distinct enough to be considered separate entities. From a diagnostic point of view, morphological details and immunohistochemistry for a
selected set of markers almost always allow for a
proper classification of a given lymphoma into
the group of LPHL or cHL, the latter of which
can be further subdivided into nodular sclerosis
cHL (NSCHL), mixed cellularity cHL (MCCHL),
lymphocyte-depleted cHL (LDCHL), and lymphocyte-rich cHL (LRCHL) [1].
The following sections summarize the key
morphological aspects and important immunohistochemical features of HL, as well as key biological and genetic features of the HL tumor
cells. For microenvironmental, clinical, and epidemiologic parameters, please refer to the respective other chapters of this book.
Although the morphology of the tumor cell population of LPHL can occasionally mimic
Hodgkin and Reed-Sternberg (HRS) cells of
cHL, in most instances the tumor cells in LPHL,
which are termed lymphocyte-predominant (LP)
cells according to the current WHO classification (previously called L&H cells, for lymphocytic and/or histiocytic Reed-Sternberg (RS)
cell variants), carry one large nucleus that is
often multilobated (“popcorn cell”) (Fig. 3.1a).
In contrast to classic HRS cells, the number of
nucleoli is increased, but they are usually less
prominent and less eosinophilic. LP cells are
found in a nodular or follicular background that
is dominated by small B lymphocytes that usually express IgD, but a more diffuse growth pattern can also be encountered, especially during
progression. The follicular infiltration pattern is
highlighted by the presence of CD21-positive
follicular dendritic cells that tend to form a welldeveloped meshwork in the nodules.
Immunohistochemically, LP cells demonstrate a
complete B cell phenotype with expression of
CD20, CD75, and, frequently, CD79a (Fig. 3.1b;
Table 3.1). Moreover, the essential B cell transcription factors BOB.1 and OCT-2 are usually
positive, and the expression of BCL6 and activation-induced cytidine deaminase (AID) is well
in line with a germinal center (GC) derivation of
the tumor cells, although CD10 is generally negative [1–3]. The negativity of the tumor cells for
CD30, CD15, and Epstein-Barr virus (EBV)
helps to distinguish LP cells from HRS cells in
Fig. 3.1 Nodular lymphocyte-predominant Hodgkin
lymphoma (LPHL). (a) HE-stained lymph node infiltrate
showing multiple characteristic, multilobated tumor
cells—termed lymphocyte-predominant (LP) cells—in a
background of small lymphocytes and histiocytes (×400).
(b) Strong CD20 expression in LP cells, but also in reactive, small B cells in the background (×400). Note that
some of the tumor cells show rosetting by a CD20-­
negative lymphocyte population. These cells are T cells
that often express the follicular T helper cell marker PD-1
3.1.1
Nodular Lymphocyte-­
Predominant Hodgkin
Lymphoma
3
Pathology and Molecular Pathology of Hodgkin Lymphoma
49
Table 3.1 Genetic and phenotypic features of HRS and LP cells
Feature
Phenotype
CD30 expression
CD15 expression
B cell receptor expression
Loss of most B cell markers
Expression of germinal center (GC) B cell markers (e.g.,
BCL6, activation-­induced cytidine deaminase (AID))
Expression of markers for non-B cells (e.g., CD3,
granzyme B, CCL17)
Putative cell of origin
EBV positivity
Signaling pathways
NF-κB activation
JAK/STAT activation
Aberrant expression of multiple RTKs
PI3K/AKT activation
AP-1 activation
Genetic lesions
NFKBIA mutations
NFKBIE mutations
TNFAIP3 mutations
REL gains/amplifications
MAP3K14 (NIK) gains/amplifications
BCL6 translocations
JAK2, PD-L1, PD-L2, JMJD2C gains/amplification
STAT6 mutations, gains
SOCS1 mutations
PTPN1 mutations
GNA13 mutations
ITPKB mutations
XPO1 mutations
B2M mutations
MHC2TA translocations
SGK1
DUSP2
JUNB
HRS cells
LP cells
Yes
Yes (~70%)a
No
Yes
Rarely
Rare
No
Yes
Modest
Yes
Frequently
No
Defective, pre-apoptotic
germinal center B cell
Yes (~40%)
Germinal center B
cell
No
Yes
Yes
Yes (~60–100%)
Yes
Yes
Yes
Yes
Yes (~40%)
n.a.
Partly
Yes (~10–20%)
Yes (~10%)
Yes (~40%)
Yes (~50%)
Yes (~25%)
Rare
Yes (~30%)
Yes (~30%)
Yes (~40%)
Yes (~20%)
Yes (~20%)
Yes (~15%)
Yes (~20%)
Yes (~30%)
Yes (~15%)
n.a.b
n.a.b
n.a.b
No
n.a.
No
No
n.a.
Yes (~50%)
No
n.a.
Yes (~50%)
n.a.
n.a.
n.a.
n.a.
n.a.
n.a.
Yes (~50%)
Yes (~50%)
Yes (~50%)
n.a. not analyzed, RTK receptor tyrosine kinase
Numbers in brackets refer to the percentage of positive cases
b
No mutations reported in 2 whole exome sequencing studies of together 44 cases of cHL and an exome sequencing
analysis of 6 cHL cell lines [8–10]
a
cHL, although occasionally a weak positivity for
CD30 can be present in LP cells (Table 3.1).
Whereas in initial lesions small B cells dominate
the background, histiocytes and T cells may
become more prominent during the evolution of
LPHL, to an extent that LPHL may be hardly
distinguishable from T cell/histiocyte-­rich large
B cell lymphoma (THRLBCL). “Variant histol-
ogy” (e.g., depletion of small B cells in the
background or unusual localization of the LP
cells) appears to be associated with an inferior
prognosis [4]. A prominent feature of LPHL is
the often impressive rosetting of LP cells by T
cells that belong to the subset of follicular T
helper cells and therefore express CD57 and
PD-1 [5–7].
A. Rosenwald and R. Küppers
50
3.1.2
lassical Hodgkin Lymphoma:
C
The HRS Cells
occasionally with prominent staining of the Golgi
area of the tumor cell. However, CD15 is negative in a significant proportion of cHL (20–25%)
The characteristic tumor cell of cHL, the RS cell, and therefore not required to establish the diagis large and contains at least two nuclear lobes or nosis of cHL [1]. CD45 is usually negative, as are
nuclei, usually with a prominent nuclear mem- the B cell transcription factors BOB.1 and OCT-­
brane (Fig. 3.2a). In contrast to LP cells in LPHL, 2. In the vast majority of cases, the derivation of
the nucleoli of RS cells are often eosinophilic. the tumor cells from the B cell lineage is indiThe mononuclear variant of RS cells is termed cated by a nuclear positivity for the B cell-­specific
the Hodgkin cell. However, the morphological activator protein PAX5/BSAP, but the staining is
spectrum of the tumor cell population in cHL can usually weaker compared to the staining intensity
be broad and includes variants such as lacunar in the small reactive B cell population in the
cells and mummified cells. In general, the tumor background of the infiltrate [11]. CD20 exprescells in cHL are called Hodgkin and Reed-­ sion can be observed in HRS cells in 30–40% of
Sternberg cells. Immunohistochemically, the cases, but the expression is frequently restricted
HRS cells stain positive for CD30 (Fig. 3.2c), to a subset of the tumor cell population, and even
and CD15 is coexpressed in the majority of cases, within one HRS cell, it is of varying intensity in
Fig. 3.2 Classical Hodgkin lymphoma (cHL). (a)
Characteristic Hodgkin and Reed-Sternberg (HRS) cells
in a mixed background of small lymphocytes, histiocytes,
and eosinophils in a mixed cellularity cHL (MCCHL)
(HE, ×400). (b) Nodular sclerosis subtype of cHL that
demonstrates thick collagen bands surrounding the nodular infiltrates (PAS, ×20). (c) CD30 expression in HRS
cells (×400). (d) Immunohistochemical staining for latent
membrane protein 1 (LMP1) shows Epstein-Barr virus
(EBV) association of HRS cells (×400)
3
Pathology and Molecular Pathology of Hodgkin Lymphoma
different parts of the cell membrane. In comparison to CD20 expression, CD79a expression is
observed less frequently [12, 13]. An EBV association, either demonstrated by immunohistochemical staining for LMP1 (latent membrane
protein 1; Fig. 3.2d) or by EBER in situ hybridization, is found in a significant proportion of
cHL, but the frequency varies considerably
between different histological subtypes and
across geographical areas [1]. Whether cHL
cases exist with a bona fide derivation from the T
cell lineage is currently a matter of debate. Single
cases have been reported, in which a T cell receptor rearrangement could be proven in the HRS
cells [14, 15], but others argue that such cases
might represent only mimics of cHL which are
not to be included in a disease entity that—based
on fundamental principles of current lymphoma
classification schemes—is of B cell derivation
[16]. HRS cells reside in a cellular background
that varies among the different histological subtypes of cHL which will be discussed in the following sections.
3.1.2.1 N
odular Sclerosis Classical
Hodgkin Lymphoma
In NSCHL, affected lymph nodes frequently
show a markedly thickened capsule and a nodular infiltrate whereby individual nodules are surrounded by broad collagen bands (Fig. 3.2b).
HRS cells are present in a background of small
lymphocytes and other nonneoplastic cells such
as histiocytes and eosinophils. The number of
HRS cells can vary significantly between
NSCHL cases and also within a single infiltrated
lymph node. Occasionally, HRS cells can form
sheets that can be associated with necrosis and
an
intense
fibrohistiocytic
reaction.
Morphologically, HRS cells in NSCHL often
show a retraction artifact of the cytoplasmic
membrane that appears to be a consequence of
formalin fixation, which has led to the term
“lacunar cell variant” of HRS cells. The immunohistochemical phenotype of HRS cells in
NSCHL as described above is the classic phenotype; however, association with EBV is less
common as compared to other cHL subtypes,
especially MCCHL.
51
3.1.2.2 M
ixed Cellularity Classical
Hodgkin Lymphoma
HRS cells in MCCHL usually have a classic morphological appearance and are scattered in a
background that can contain small lymphocytes,
eosinophils, neutrophils, plasma cells, and histiocytes. The infiltration pattern can be diffuse or
vaguely nodular; sometimes, the lymph node
architecture and especially some B cell areas are
partially preserved leading to an interfollicular
infiltration pattern. The characteristic features of
other histologic cHL subtypes (e.g., the formation of nodular collagen bands) are absent, and,
thus, MCCHL is sometimes considered as the
“wastebasket” of cHL. The EBV association of
HRS cells is the highest among all cHL subtypes
and can reach 75% [1].
3.1.2.3 Lymphocyte-Depleted Classical
Hodgkin Lymphoma
LDCHL is the rarest histological subtype of cHL
(<1% of cases) and probably the most problematic one to define. It is characterized by an
increased number of HRS cells present in the
infiltrate and/or depletion of small lymphocytes
in the nonneoplastic background population. In
some cases, HRS cells are of anaplastic appearance, and in other cases, the background is composed of extensive diffuse fibrosis. However, if
the pattern of fibrosis is nodular and therefore
characteristic of NSCHL, a given case should be
classified as NSCHL, regardless of whether there
is a high number of HRS cells. Since the definition of LDCHL has changed over the past
decades, some of the established clinical and biological features appear outdated in the context of
the current definition. Moreover, with the increase
in knowledge and the development of additional
immunohistochemical markers, some of the cHL
cases that were previously assigned to the LDCHL
category would nowadays be included into borderline categories or even different entities [1].
3.1.2.4 Lymphocyte-Rich Classical
Hodgkin Lymphoma
In LRCHL, the HRS cells are present in a
lymphocyte-­rich background that can be nodular
or, rarely, diffuse. Often, B cell follicles are
A. Rosenwald and R. Küppers
52
partially preserved with recognizable GC, and
HRS cells can be found in expanded mantle and
marginal zones, thus providing a B cell-rich
background. HRS cells in LRCHL may resemble
LP cells in LPHL morphologically to such an
extent that they are indistinguishable from each
other without additional immunohistochemical
characterization. It is of significance that eosinophils and neutrophils should be absent from the
nodular infiltrates and may only be found in low
numbers in interfollicular zones and close to vascular structures. The immunophenotype of the
HRS cells is classic, and an EBV association is
occasionally observed, though at a lower fre­
quency compared to MCCHL [1].
3.2
Differential Diagnosis
In most instances, the diagnosis of LPHL and
cHL is unambiguous on the basis of morphological, clinical, and, especially, immunohistochemical features (Table 3.1). However, a gray area
between cHL and diffuse large B cell lymphoma
(DLBCL), specifically with primary mediastinal
large B cell lymphoma (PMBL), has long been
known, and the most recent WHO classification
introduced the category of “B cell lymphoma,
unclassifiable, with features intermediate
between DLBCL and classical Hodgkin lymphoma” [1]. It is important to note that lymphomas falling into this category are not considered
a separate disease entity; rather, it was felt that
lymphomas in which there is a discordance
between morphological aspects of the infiltrate
and the expected immunophenotype should be
labeled as “intermediate” to allow a more precise
definition of biological and clinical features of
these lymphomas in the future. Frequently, these
borderline lymphomas present with large mediastinal masses. Morphologically, they consist of
large, pleomorphic B cells that grow in a sheetlike pattern in a background of a fibrotic stroma.
A subset of the tumor cells may resemble HRS
cells, specifically the lacunar variant, and parts of
the infiltrate may correspond to the growth pattern of cHL, particularly the nodular sclerosis
subtype. Immunophenotypically, there is often a
preserved expression program of cHL including
expression of CD30 and CD15, while markers of
the B cell lineage that are often downregulated in
cHL, such as CD20 and CD79a, are equally
expressed in the tumor cells [1]. It is important to
note that these gray zone lymphomas appear to
be more common in male patients, in contrast to
NSCHL and PMBL that are more frequent in
females [17]. Clinically, these tumors may
behave more aggressively than NSCHL and
PMBL; it has to be determined in the future
whether treatment regimens for aggressive B cell
lymphomas or for cHL are more beneficial.
The differential diagnosis between cHL and
ALK-negative anaplastic large cell lymphoma
(ALCL) of T cell lineage can usually be resolved
using an appropriate panel of immunohistochemical markers including T cell, cytotoxic, and other
markers. Problems arise when morphological
features favor cHL, but tumor cells lack PAX5/
BSAP expression while cytotoxic markers are
expressed. As discussed above, it is a matter of
current debate whether such cases should be
grouped into the cHL category or diagnosed as
ALCL. Remarkably, a global gene expression
study revealed surprisingly few consistent differences in the gene expression of HRS cells and
ALK-negative ALCL cells [18].
Finally, EBV-associated lymphoproliferations, e.g., in the context of a coexisting T cell
non-HL as well as EBV-associated DLBCL of
the elderly, a subgroup of DLBCL introduced in
the new WHO classification [1], can harbor HRS
or HRS-like cells and therefore mimic cHL [19].
Besides other morphological and immunohistochemical features and information on the clinical
setting, the pattern of EBV infection, determined
by LMP1 staining or EBER in situ hybridization,
might help to distinguish between these tumors.
3.3
istogenesis of HRS and LP
H
Cells
3.3.1
ellular Origin of HRS and LP
C
Cells
The unusual immunophenotype of HRS cells,
which does not resemble any normal hematopoietic cell, has hampered the identification of the
3
Pathology and Molecular Pathology of Hodgkin Lymphoma
cellular origin of these cells considerably.
Moreover, only few cell lines were available for
detailed genetic studies, and the rarity of the HRS
cells in the tissue posed a problem for their
molecular analysis. Finally, by microdissection
of HRS cells from tissue sections and single-cell
polymerase chain reaction analysis of these cells,
it was clarified that HRS cells derive from B cells
in nearly all cases [20, 21]. This is because rearranged immunoglobulin (Ig) heavy (IgH) and
light (IgL) chain gene rearrangements were
detected in these cells. The detection of identical
IgV gene rearrangements in the HRS cells of a
given HL case also established the monoclonal
nature of these cells, a hallmark of malignant
cancer cells. With a few exceptions, somatic
mutations were detected in the rearranged V
genes of HRS cells [20–23]. As the process of
somatic hypermutation, which generates such
mutations, is specifically active in antigen-­
activated mature B cells proliferating in the GC
microenvironment in the course of T-dependent
immune responses [24], the presence of mutated
IgV genes in the HRS cells established their derivation from GC-experienced B cells. A surprising finding was that about 25% of cases of cHL
showed destructive IgV gene mutations, such as
nonsense mutations or deletions causing frameshifts that rendered originally functional V region
genes nonfunctional [20]. When such mutations
happen in normal GC B cells, these cells quickly
undergo apoptosis. On this basis, it was proposed
that HRS cells in these cases derive from pre-­
apoptotic GC B cells that were rescued from
apoptosis because they harbored or acquired
some transforming events [20, 25]. It is important
to note that crippling mutations, such as those
generating premature stop codons, represent only
a small fraction of disadvantageous IgV gene
mutations that cause apoptotic death of GC B
cells, and it is therefore likely that also most or
even all other cases of cHL are derived from pre-­
apoptotic GC B cells. Even a few HL with unmutated IgV genes may derive from these precursors,
because GC founder cells proliferating in GC
become prone to apoptosis before the onset of
somatic hypermutation activity [26]. The GC B
cell origin of HRS cells was further supported by
the molecular analysis of composite lymphomas,
53
composed of a cHL and a B cell non-HL. Such
cases are often clonally related and show an
intriguing pattern of shared as well as distinct
somatic V gene mutations [27–30]. This pattern
supports the assumption that both lymphomas
were derived from distinct members of a proliferating GC B cell clone.
A comparison of the transcriptomes of HRS
cells and normal GC and extrafollicular CD30+ B
cells revealed that HRS cells are in their global
gene expression pattern more similar to the normal CD30+ B cells than to bulk GC B cells [31].
However, a direct derivation of HRS cells from
CD30+ GC B cells seems unlikely, as CD30+ GC
B cells are positively selected GC B cells with
functional BCR that are preparing to return to the
dark zone of the GC for a further round of proliferation and IgV gene mutation. Perhaps, in the
course of their malignant transformation, the
HRS cell precursors that managed to escape from
apoptosis acquired the gene expression program
of the positively selected and proliferation prepared CD30+ GC B cells.
A few cases of cHL appear to originate from T
cells, because T cell receptor gene rearrangements were detected in some cases diagnosed as
HL and expressing some typical T cell molecules
[14, 15]. However, it is debated whether these are
true HL (see above). Remarkably, among HL
cases with expression of one or more T cell markers, the majority nevertheless derives from B
cells [14, 15].
The expression of multiple B cell markers by
LP cells of LPHL already indicated a B cell derivation of these cells. Moreover, LP cells express
several markers typically expressed by GC B
cells, such as BCL6, AID, centerin, and hGAL,
and the cells grow in a follicular pattern in close
association with typical constituents of normal
GC, i.e., follicular dendritic cells and GC-type T
helper cells [2, 3, 5, 6, 32, 33]. This pointed to a
close relationship between LP cells and GC B
cells. This is indeed supported by the detection of
clonally related and somatically mutated IgV
genes in these cells [21, 34–36]. As opposed to
cHL, the V genes are selected for functionality,
and a fraction of cases shows ongoing somatic
hypermutation during clonal expansion, a hallmark of GC B cells [21, 34, 35]. Thus, these
A. Rosenwald and R. Küppers
54
findings altogether indicate a GC B cell origin of
LP cells. A large-scale gene expression profiling
of isolated LP cells in comparison to the main
subsets of mature B cells has led to a further
specification of the derivation of LP cells by
showing that the gene expression pattern of LP
cells resembles that of GC B cells that have
already acquired some features of post-GC memory B cells [37].
3.3.2
elationship of Hodgkin Cells
R
and Reed-Sternberg Cells
and Putative HRS Cell
Precursors
The relationship of the mononucleated Hodgkin
cells to the multinuclear RS cells and the potential existence of HRS precursor cells has been a
matter of debate. Based on the “mixed” phenotype of HRS cells and many numerical chromosomal aberrations in these cells, it has been
speculated that HRS cells as such or, specifically,
the RS cells may derive from cell fusions of different cells (e.g., a B cell and a non-B cell).
However, a detailed study of antigen receptor loci
revealed that HRS cells do not carry more than
two different alleles of these loci, which strongly
supports the assumption that these cells do not
derive from cell fusions [38]. Several studies of
HL cell lines showed that the mononuclear
Hodgkin cells give rise to the RS cells and that
the latter have little proliferative activity [39–41].
Long-term time lapse-microscopy analyses
revealed that mononucleated Hodgkin cells
undergo incomplete cytokinesis and refusion to
give rise to the multinucleated RS cells [42, 43].
Two studies reported the existence of a small
subpopulation of side population cells among the
mononuclear Hodgkin cells. Side population
cells extrude the Hoechst dye, because they
express multidrug transporters, such as MDR1
and/or ABCG2. In several types of cancers, there
is an overlap between side population cells and
cancer stem cells. Side population cells of cHL
cell lines were CD30+CD20- and showed
increased resistance against chemotherapeutic
drugs [44, 45]. However, it has not yet been
determined whether they have a higher capacity
to sustain the HRS cell clone in long term than
other mononuclear Hodgkin cells, and the fact
that side population cells were not identified in
all cHL cell lines analyzed argues against an
essential role of these cells for the survival of the
HRS cell clone.
Another debated issue relates to the question
whether the CD30+ typical HRS cells represent
the entire tumor clone in HL or whether members
of the HRS cell clones exist among small CD30–
cells. An initial study for numerical chromosomal
abnormalities indeed suggested that such CD30–
clone members might exist [46]. However, trisomies of chromosomes as studied in that work are
not a stringent clonal marker. Moreover, a molecular analysis of EBV-positive HL cases for members of the malignant clones among small,
CD30– EBV+ B cells in the HL lymph nodes suggested that the small EBV+ B cells rarely, if at all,
belong to the HRS cell clones [47]. Two HL cell
lines were reported to contain small subpopulations of CD20+CD30–Ig+ B cells coexpressing the
stem cell marker aldehyde dehydrogenase
(ALDH) [48]. These cells had clonogenic potential and gave rise to the typical HRS cells of these
lines. It is important to note that ALDHhigh cells
were also detectable in the peripheral blood of
most HL patients, and it was reported that these
cells were often clonally related to the HRS cells
[48]. However, the clonal relationship between
the HRS cells and ALDHhigh peripheral blood B
cells was not clearly shown [49], so it remains to
be clarified whether ALDHhigh B cells indeed represent precursors of the HRS cell clones. A previous study using a highly sensitive PCR for HRS
cell-specific Ig gene rearrangements failed to
detect members of the HRS cell clone in the
peripheral blood or bone marrow of two HL
patients [50].
3.4
Genetic Lesions
HRS cells have a much higher number of chromosomal aberrations, including multiple numerical as well as structural abnormalities, than most
other lymphomas [51]. However, it is still unclear
3
Pathology and Molecular Pathology of Hodgkin Lymphoma
whether this is mostly a side effect of some type
of genetic instability and whether the expression
of specific oncogenes or tumor suppressor genes
is recurrently affected by these lesions. When the
B cell origin of HRS cells became clear, HRS
cells were studied for the presence of chromosomal translocations involving the Ig loci, as
such translocations are a hallmark of many B cell
lymphomas. Fluorescence in situ hybridization
(FISH) studies indeed provided evidence for such
translocations in about 20% of cases, but most of
the translocation partners involved remain to be
identified [52, 53]. In a few cases, the translocation partners were BCL2, BCL3, REL, BCL6, or
MYC [52–55]. Recurrent translocations affecting
the major histocompatibility complex (MHC)
class II transactivator (MHC2TA) were detected
in about 15% of cHL cases [56]. These translocations appear to cause downregulation of MHC
class II expression by HRS cells. In LPHL, translocations of the BCL6 gene have been found in
about 30% of cases [57, 58]. These translocations
can involve the Ig loci, but also multiple other
partners [59].
Due to the difficulty to analyze the few HRS
and LP cells for mutations in oncogenes and
tumor suppressor genes, only relatively few of
such genes have been analyzed so far in these
cells. There was a major interest to understand
the apoptosis resistance of HRS cells, but it
turned out that mutations in the CD95 gene, an
important death receptor, as well as in members
of the CD95 signaling pathway (FADD, caspase
8, caspase 10) were rare or not found at all [60–
62]. Likewise, no mutations were found in the
BCL2 family member BAD, and also ATM lesions
are very rare [63–65]. The TP53 tumor suppressor gene was mutated in less than 10% of cases
where the exons of TP53 usually carrying mutations were studied in isolated HRS cells [66, 67].
However, studies of HL cell lines indicate that
HRS cells may additionally carry untypical TP53
mutations and that the frequency of TP53 mutations may therefore be higher than previously
thought [68]. MDM2, a negative regulator of
TP53, frequently shows gains in HRS cells,
which might contribute to impaired functions of
TP53 in these cells [69].
55
Further candidate gene mutation studies
revealed frequent mutations in the exportin 1
gene (XPO1) [70], which encodes a nuclear
expert receptor for numerous RNAs and proteins,
and inactivating mutations in and deletions of
CD58 [71, 72]. CD58 is important for targeting
of cells by cytotoxic T cells and NK cells, so that
CD58 inactivation may contribute to immune
escape of HRS cells from an attack by these cells.
HRS cells show constitutive activity of the
NF-κB transcription factor (see below), which is
essential for the survival of these cells. The
mechanisms of this activation were originally not
understood. Consequently, members and regulators of this signaling pathway were studied for
genetic lesions (Table 3.1). Inactivating mutations in the main NF-κB inhibitor NFKBIA
(IκBα) were found in about 10–20% of HL cases
and also in several HL cell lines (Fig. 3.3) [73–
76]. Recurrent mutations were also detected in
another NF-κB inhibitor, NFKBIE (IκBε) [77,
78]. Inactivating mutations or deletions in two
further negative regulators of NF-κB signaling,
CYLD and TRAF3, have also been detected in
HL cell lines and a few primary cases, but overall
these events are rare [79, 80]. Moreover, HRS
cells frequently harbor genomic gains or amplifications of the REL gene [81–83], encoding an
NF-κB family member, and a correlation between
such gains and strong REL protein expression
was found [84]. The MAP3K14 gene, which
encodes the NIK kinase, a major activating component of the alternative NF-κB pathway, shows
gains or amplifications in about 15% of cHL [79,
85]. Also the IκB family member BCL3, which
acts as a positive regulator of NF-κB activity, is
affected by chromosomal gains or translocations
in a small fraction of cHL [86, 87]. Somatic and
clonal inactivating mutations were found in the
TNFAIP3 gene in about 40% of cHL [88, 89].
TNFAIP3 encodes for the A20 protein, which is a
dual ubiquitinase and deubiquitinase that functions as a negative regulator of NF-κB. It inhibits
signaling from the receptor-interacting protein
(RIP) and TNF receptor-associated factors
(TRAFs) to the IKK kinases, which are essential
mediators of NF-κB signaling. TNFAIP3 mutations were mainly found in EBV-negative cases.
A. Rosenwald and R. Küppers
56
CD40
EBV
infection
(40%)
LMP1
RANK
BCMA
CD40
TACI
cytokine
receptor
CD30
RIP TRAF
TNFAIP3
mutations
(40%)
TNFAIP3
NEMO
Classical
NF-κB
pathway
IKKα
NFKBIA and
NFKBIE mutations
(10-20%)
Alternative
NF-κB
pathway
JAK2 gains
(30%)
PTPN1
IKKα
IKKβ
JAK
IKKα
SOCS1
PTPN1
mutations
(20%)
STAT STAT
IκBα/
IκBε
p50
REL
amplification
(40%)
p50
NIK
NIK gains
(25%)
p65
p100 RELB
Proteasomal
degradation
BCL3 gains or
translocations
(rare)
p65
JAK/STAT
pathway
SOCS1
mutations
(40%)
STAT6
mutations
& gains
(30%)
cytoplasm
p52 RELB
nucleus
BCL3
p50 p50
Fig. 3.3 NF-κB and JAK/STAT activity in HRS cells. In
the classical NF-κB signaling pathway, stimulation of
numerous receptors leads via TNF receptor-associated
factors (TRAFs), which are often associated with the
receptor-interacting protein (RIP), to activation of the
IKK complex, which is composed of IKKα, IKKβ, and
NEMO. The IKK complex subsequently phosphorylates
the NF-κB inhibitors IκBα and IκBε. This marks them for
ubiquitination and subsequent proteasomal degradation.
Thereby the NF-κB transcription factors (p50/p65 or p50/
REL heterodimers) are no longer retained in the cytoplasm and translocate into the nucleus, where they activate multiple genes. The signal transduction from TRAFs/
RIP to the IKK complex can be inhibited by TNFAIP3,
which removes activating ubiquitins from RIP and TRAFs
and additionally links ubiquitins to these molecules to
mark them for proteasomal degradation. In the alternative
NF-κB pathway, activation of receptors such as CD40,
BCMA, and TACI causes stimulation of the kinase NIK,
which then activates an IKKα complex. Activated IKKα
processes p100 precursors to p52 molecules, which translocate as active p52/RELB NF-κB heterodimers into the
nucleus. HRS cells show constitutive activity of the classical and alternative NF-κB signaling pathway. This activity is probably mediated by diverse mechanisms, including
receptor signaling through CD40, RANK, BCMA, and
TACI; genomic REL and MAP3K14 (NIK) amplification;
STAT STAT
destructive mutations in the TNFAIP3, NFKBIA, and
NFKBIE genes; and signaling through the EBV-encoded
LMP1. The role of CD30 signaling in HRS cells is controversially discussed. HRS cells may also harbor nuclear
BCL3/(p50)2 complexes, and in a few cases, the strong
BCL3 expression appears to be mediated by genomic
gains or chromosomal translocations. The JAK/STAT
pathway is the main signaling pathway for cytokines.
Upon binding of cytokines to their receptors, members of
the JAK kinase family become activated by phosphorylation. The activated JAKs then phosphorylate and thereby
activate STAT transcription factors. These phosphorylated
factors homo- or heterodimerize and translocate into the
nucleus where they activate target genes. Main inhibitors
of the JAK/STAT pathway are the phosphatase PTPN1
and SOCS (suppressor of cytokine signaling) factors,
which function by binding to JAK molecules and inhibiting their enzymatic activity and additionally by inducing
proteasomal JAK degradation. In HRS cells, STAT3, 5,
and 6 are constitutively active. Besides activation of cytokine receptors (e.g., IL13 receptor and IL21 receptor)
through cytokines, activation of this pathway is mediated
by genomic gains or rare translocations of the JAK2 gene,
activating mutations in the STAT6 gene, and frequent inactivating mutations in the SOCS1 and PTPN1 gene. The
frequency of genetic lesions and viral infections affecting
NF-κB or STAT activity in cHL cases is indicated
3
Pathology and Molecular Pathology of Hodgkin Lymphoma
Nearly 70% of EBV– cases carried TNFAIP3
mutations, indicating that EBV infection and
A20 inactivation are alternative pathogenetic
mechanisms in HL [88, 89]. As LMP1 of EBV,
which is expressed in EBV-positive HRS cells,
mimics an active CD40 receptor and signals
through NF-κB [90, 91], LMP1 may replace the
role of A20 inactivation in EBV+ HL.
As it was recently revealed that also the LP cells
of LPHL show strong constitutive NF-κB activity
[37], also these cells were studied for mutations in
NFKBIA and TNFAIP3, but clonal destructive
mutations were not found (Table 3.1) [92].
Genetic lesions were also found in members
of the JAK/STAT pathway, which is constitutively activated in HRS and LP cells. In about
40% of cases analyzed, both HRS and LP cells
showed somatic mutations in the SOCS1 gene,
which encodes a main inhibitor of STAT signaling (Fig. 3.3) [93, 94]. In HRS cells, recurrent
mutations were additionally found in the gene of
another negative regulator of JAK/STAT signaling, namely, the PTPN1 gene, which encodes a
phosphatase [95]. Furthermore, a fraction of cHL
cases show genomic gains or amplifications of
the JAK2 locus, which encodes one of the kinases
activating the STAT factors (Table 3.1) [82, 96].
Importantly, the genomic gains at 9p24 do not
only affect the JAK2 locus, but additionally the
PD-L1, PD-L2, and JMJD2C genes [97, 98].
PD-L1 and PD-L2 are inhibitory receptors for
PD1-positive T cells and may hence inhibit a
cytotoxic T cell attack on HRS cells. JMJD2C
encodes a histone demethylase and plays a role in
the epigenetic remodeling of HRS cells. Finally,
the JAK2 gene is in rare instances also deregulated by chromosomal translocations [99].
Activating point mutations in the STAT6 gene and
genomic gains involving this gene were also
detected in HRS cells [10, 100]. Thus, multiple
types of genetic lesions cause a constitutive JAK/
STAT signaling, suggesting an essential role of
its deregulated activity for cHL pathogenesis.
With the availability of high-throughput
sequencing methods, tumor cells can now be studied for genetic lesions at a genome-wide level. An
exome sequencing analysis of six cHL lines and
the only LPHL cell line (DEV) revealed over 400
57
genes mutated in at least two of the lines [8]. This
is a valuable database that should be considered
when performing functional studies with these cell
lines. A first whole exome sequencing study of primary HRS cells used flow-cytometry isolated lymphoma cell from ten cases of cHL [9]. Between 100
and 500 somatic mutations were found per case. A
main finding of this analysis was recurrent inactivating mutations in the B2M gene. B2M is essential
for MHC class I expression, so that the loss of its
expression presumably leads to immune evasion
from CD8+ cytotoxic T cells. Other novel recurrent
mutations identified in that work affect several histone genes, the inositol-trisphosphate 3-kinase B
(ITPKB), the B cell transcription factor EBF1, and
the G protein subunit GNA13 [9]. Tiacci and colleagues performed a whole exome sequencing
analysis of pools of HRS cells microdissected from
34 cases of cHL [10]. A median of 47 non-­silent
somatic mutations in the exomes was found. This
study confirmed recurrent mutations in ITPKB and
GNA13, and newly revealed recurrent point mutations in STAT6, further adding to the complexity of
JAK/STAT deregulation in cHL. Although only
four EBV+ cases were included in the study by
Tiacci et al., it seems that such cases carry considerably fewer somatic mutations than the EBVnegative cases. A mutation study of LP cells of
LPHL was based on a whole genome analysis of
DLBCL clonally related to LPHL in the same
patient, followed by targeted sequencing analysis
of microdissected LP cells. In this work, three
genes were found to be each mutated in about half
of the cases of LPHL (also in cases without cooccurring DLBCL), namely, the genes encoding the kinase SGK1, the AP-1 family member
JUNB, and the phosphatase DUSP2 [101].
3.5
Deregulated Transcription
Factor Networks
and Signaling Pathways
3.5.1
The Lost B Cell Phenotype
Early immunohistochemical studies already
revealed that HRS cells usually do not express
typical B cell markers, such as CD20, CD79b, or
58
the BCR [13, 102–104]. This lack of expression
of B cell markers was indeed one of the reasons
why the B cell origin of HRS cells was not
revealed until genetic studies for Ig gene rearrangements unequivocally demonstrated a B cell
identity of these cells (see above). Gene expression profiling studies of HRS cells in comparison
to normal B cells then showed that there is a
global loss of the B cell typical gene expression
in HRS cells [105]. This downregulation involved
all types of genes with important functions in
these cells, for example, cell surface receptors
(CD37, CD53), components of signaling pathways (SYK, BLK, SLP-65), and transcription
factors (PU.B, A-MYB, SPI-B). As plasma cells
also show a downregulation of many B cell-­
typical genes, it had been speculated that HRS
cells lost their B cell gene expression and
acquired a partial plasma cell differentiation program [2, 106]. However, a gene expression profiling study of microdissected HRS cells revealed
that HRS cells have not acquired a plasma cell
phenotype [107].
Remarkably, HRS cells have retained expression of molecules that are involved in antigen-­
presenting functions and the interaction with
CD4+ T helper cells. HRS cells usually express
CD40, CD80, and CD86 and often MHC class II
[105, 108]. This indicates that an interaction
with T helper cells is important for HRS cell survival. In line with this view, HRS cells are typically surrounded by CD40L expressing CD4+ T
cells [109].
We are now beginning to understand which
factors contribute to the lost B cell phenotype of
HRS cells. First, several transcription factors that
positively regulate the expression of multiple
genes in B cells are downregulated, including
OCT-2, PU.1, EBF1, ETS1, and BOB.1 [102,
103, 110–112]. The downregulation of ETS1
may often be due to heterozygous deletions of the
gene, which have been observed in over 60% of
cHL analyzed [112]. Second, although E2A, a
master regulator of the B cell transcription program, is still expressed, HRS cells also show
deregulated expression of ID2 and ABF1 [113–
115], which bind to E2A and inhibit its function
[114]. The physiological role of ABF1 is poorly
A. Rosenwald and R. Küppers
understood, but ID2 is normally expressed in
dendritic cells and natural killer cells, and supports the generation of these cells concomitant
with suppression of B cell development [116,
117]. Third, HRS cells express activated
NOTCH1, which normally induces T cell differentiation in lymphocyte precursors and suppresses a B lineage differentiation of such cells
[118, 119]. Activation of NOTCH1 is probably
caused by interaction with its ligand Jagged-1,
which is expressed by other cells in the HL
microenvironment [119], and by high-level
expression of the NOTCH coactivator
mastermind-­like 2 (MAML2) [120]. Moreover,
HRS cells have downregulated the NOTCH1
inhibitor Deltex1 [118]. Fourth, STAT5A and
STAT5B are activated in HRS cells and have
been reported to induce an HRS cell-like phenotype in normal B cells [121]. Constitutive active
STAT5 induced expression of CD30 and of the T
cell transcription factor GATA3 in the B cells and
led to downregulation of BCR expression.
Aberrant GATA3 expression in HRS cells is furthermore mediated by NOTCH1 and NF-κB
activity in HRS cells [122]. Fifth, the downregulation of multiple B cell genes in HRS cells is
further caused by epigenetic mechanisms, as
DNA methylation has been detected for numerous such genes [123–125]. Sixth, HRS cells
express several transcription factors that have
important roles in hematopoietic stem cells and
early lymphoid precursors, including GATA2,
BMI1, RING1, and RYBP [126–129]. The
expression of these factors may contribute to a
“dedifferentiated” phenotype of HRS cells.
Surprisingly, PAX5, the main B lineage commitment and maintenance factor, is still expressed
in HRS cells, albeit at reduced levels [11]. As
many of its direct target genes are not expressed,
it is likely that PAX5 activity is inhibited.
NOTCH1 is a candidate for this inhibition [118].
It may also be that PAX5 target genes are not
expressed because other transcription factors
needed for the efficient expression of these genes
are missing.
Expression of the myeloid specific colony-­
stimulating factor 1 receptor (CSF1R) by HRS
cells is a further important example of aberrant
3
Pathology and Molecular Pathology of Hodgkin Lymphoma
expression of a non-B cell gene in HRS cells
[130]. CSF1R expression promotes HRS cell
survival. The mechanism of its deregulated
expression is remarkable, because this is mediated by derepression of an endogenous long terminal repeat upstream of the CSF1R gene that
replaces the function of the normal CSF1R promoter [130].
The downregulation of many B cell transcription factors that also suppress the expression of
non-B lineage genes, combined with the upregulated expression of genes promoting expression
of genes of other hematopoietic cell types (e.g.,
NOTCH1, ID2), not only explains the lost B cell
phenotype of HRS cells but also the heterogeneous expression of genes specifically expressed
by dendritic cells, T cells, or other cell types. It is
an intriguing question whether the lost B cell
phenotype of HRS cells is related to their origin
from crippled GC B cells. Perhaps, due to the
stringent selection of B cells for expression of a
functional BCR (a high-affinity one in the GC),
there is a selection in HRS cell pathogenesis
downregulating the B cell gene expression program to escape the selection forces that induce
apoptosis in GC B cells with unfavorable IgV
gene mutations. The observation that enforced
re-expression of the B cell transcription factors
PU.1, FOXO1, or E2A or the pharmacological
restoration of the B cell phenotype in HL cell
lines induces apoptosis is in line with this view
[131–134]. However, the lost B cell phenotype
could also be a side effect of so far unknown
transforming events.
3.5.2
Constitutive Activation
of Multiple Signaling
Pathways
It is obvious that tumor cells need to activate and
deregulate signaling pathways and transcription
factors that promote their survival and proliferation. Nevertheless, it is striking how many of
such pathways are constitutively activated in
HRS cells, and cHL appears to be rather unique
among lymphoid malignancies in the extent to
which multiple signaling pathways contribute to
59
the survival and expansion of HRS cells. It has
already been mentioned above that HRS cells
show constitutive NF-κB activity. This activity is
essential for HRS cell survival [135] and is most
likely not only mediated by genetic lesions (see
above) but also by signaling through receptors.
NF-κB factors of both the canonical pathway
(p50/p65) and the noncanonical NF-κB pathway
(p52/RelB) are activated (Fig. 3.3). HRS cells
express the TNF receptor family members CD30,
CD40, RANK, TACI, and BCMA, which activate
NF-κB, and cells expressing the respective
ligands are found in the HL microenvironment
[109, 136–140]. There are, however, conflicting
data about the role of CD30 in NF-κB activation
[141, 142]. In EBV-positive cases of cHL, the
virally encoded LMP1 mimics an active CD40
receptor and hence also contributes to NF-κB
activation [143].
Another central signaling pathway, which is
like NF-κB activated both by genetic lesions and
by ligand-mediated receptor triggering, is the
JAK/STAT pathway (Fig. 3.3). This is the main
signaling pathway for cytokines. Activation of
cytokine receptors causes activation of JAK
kinases which in turn phosphorylate and thereby
activate STAT transcription factors. The phosphorylated STAT factors dimerize and then translocate into the nucleus where they activate
transcription of target genes. HRS cells show
activation of STAT3, STAT5, and STAT6 [121,
144–146]. The activation of STAT6 is at least
partly mediated by signaling through IL13. As
HRS cells express IL13 and its receptor, STAT6
activation can be mediated through an autocrine
stimulation loop [147, 148]. Signaling through
the IL21 receptor contributes to STAT3 and
STAT5 activation in HRS cells, which is also
enhanced by the NF-κB activity in the cells [121,
149, 150]. As mentioned above, STAT5 activity
may contribute to the lost B cell phenotype of
HRS cells. Inhibition of STAT activity in HL cell
lines resulted in reduced proliferation of the cells,
further supporting an important pathogenetic role
of this signaling pathway [144, 145, 147].
Receptor tyrosine kinases (RTKs) are important regulators of cell growth, survival, and proliferation. In multiple cancers, specific RTKs are
60
A. Rosenwald and R. Küppers
activated, often by somatic mutations [151]. In toxic for these cells, revealing an essential role of
contrast, HRS cells show multiple activated this factor in cHL pathophysiology [162].
RTKs, and their activation does not appear to be
Finally, also the phosphatidylinositol 3-kinase
due to activating mutations but at least partly to (PI3K)/AKT pathway, which is a main promoter
ligand-mediated stimulation [152]. RTKs that are of cell survival, shows activity in HRS cells [164,
often expressed in varying combinations in HRS 165]. AKT is a serine/threonine kinase that is
cells include PDGFRA, DDR2, EPHB1, RON, activated in HRS cells, as evident from its phosTRKA, TRKB, CSF1R, and MET [130, 152, phorylated state and phosphorylation of known
153]. The expression of most of these is aberrant, target proteins [164, 165]. Inhibition of AKT in
as they are not expressed by normal GC B cells HL cell lines causes cell death, suggesting an
[130, 152]. They are also usually not expressed important role of active AKT in HRS cell surby other B cell non-HL, showing that this is a vival [164, 165]. PI3K may be activated in HRS
specific feature of HL among B cell lymphomas cells by signaling through CD30, CD40, RANK,
[152, 154]. Expression of multiple RTKs is most and RTK. Moreover, downregulation of the AKT
pronounced in EBV-negative cases of cHL, sug- inhibitor INPP5D in HRS cells may further congesting that EBV activates pathways in HRS tribute to strong AKT activity in these cells [107].
cells replacing the function of RTKs [155]. For
While we have a relatively detailed insight
PDGFRA, TRKA, and CSF1R, a growth-­ into signaling pathways active in HRS cells, less
inhibitory effect has been shown upon their inhi- is known about signaling pathways constitutively
bition in HL cell lines, giving a first indication active in LP cells of LPHL. However, LP cells
that the activity of RTKs is important for HRS also show a high constitutive activity of NF-κB
cell proliferation [130, 152, 156].
[37]. RTKs are partly also aberrantly expressed
Signaling through various receptors is medi- by these cells [152], and activation of the JAK/
ated by the mitogen-activated protein kinase STAT pathway has been observed [93].
(MAPK)/ERK pathway. In HRS cells, the serine/
In conclusion, HRS cells are characterized by
threonine kinases ERK1, ERK2, and ERK5 are the deregulated and constitutive activation of
activated [157, 158]. Inhibition of their activity multiple signaling pathways and transcription
has antiproliferative effects on HL cell lines factors that contribute to the survival and prolif[158]. Signaling through CD30, CD40, and eration of these cells. The multitude of different
RANK may contribute to the stimulation of this stimulated pathways appears to be rather unique
pathway [158].
among human B cell lymphomas. Often, these
The transcription factor AP-1 acts as homo- or pathways are activated by common mechanisms,
heterodimers of JUN, FOS, and ATF compo- and they may interact in numerous ways.
nents. In HRS cells, JUN and JUNB are overexpressed and constitutively active [159]. The
overexpression of JUNB is mediated by NF-κB 3.6
Anti-apoptotic Mechanisms
[159]. AP-1 induces many target genes and promotes proliferation of HRS cells. Target genes of With a presumed origin from pre-apoptotic GC B
AP-1 include CD30 and galectin-1, the latter of cells, it is critical to understand through which
which has immunomodulatory functions [160, mechanisms HRS cells escape from apoptosis. A
161]. HRS cells also show strong expression of number of factors contributing to HRS cell surBATF3, another member of the AP-1 transcrip- vival have already been discussed in the previous
tion factor family [162, 163]. BATF3 expression section: constitutive activity of NF-κB, STAT,
is induced by STAT3 and STAT6 in HRS cells. It PI3K, NOTCH1, AP-1, RTK, and ERK. Several
forms heterodimers with JUN and JUNB, and the specific inhibitors of the two main apoptosis
proto-oncogene MYC was identified as one of pathways deserve specific mentioning. Although
the direct BATF3 target genes [162]. Importantly, HRS cells express the CD95 death receptor of
downregulation of BATF3 in HL cell lines is the extrinsic apoptosis pathway as well as its
3
Pathology and Molecular Pathology of Hodgkin Lymphoma
activating ligand, HL cell lines are resistant to
CD95-­
mediated death induction, suggesting a
specific inhibition of this pathway [166–168]. As
mentioned above, this resistance is neither due to
mutations in the CD95 receptor itself nor in its
interaction partners FADD, caspase 8, or caspase
10. However, HRS cells show strong expression
of the CD95 inhibitor CFLAR (previously known
as cFLIP, cellular FADD-like interleukin
1b-­
converting enzyme-inhibitory protein), and
this factor impairs CD95 signaling in HRS cells
[166, 167]. Inhibition of the intrinsic (mitochondrial) apoptosis pathway is probably mediated
through strong expression of the anti-apoptotic
factors BCLXL and XIAP (X-linked inhibitor of
apoptosis) and downregulation of the pro-­
apoptotic factor BIK [107, 169, 170]. BCLXL
inhibits apoptosis at the level of the mitochondrial apoptosis induction, whereas XIAP inhibits
activity of caspases 3 and 9, which are downstream executioners of the mitochondrial apoptosis program. Although HRS cells also express
pro-apoptotic SMAC, which can inhibit XIAP,
the cells show an impaired release of SMAC
from the mitochondria into the cytoplasm [171].
As mentioned above, HRS cells express high levels of the pro-apoptotic TP53 factor, but resistance to TP53-mediated apoptosis appears to be
rarely due to inactivating mutations in the TP53
gene. An important factor for the inhibition of
TP53 activity is MDM2, which is expressed at
high levels in HRS cells [172]. The functional
role of MDM2 as a TP53 inhibitor in HRS cells is
supported by the fact that HL cell lines expressing wild-type TP53 are rendered apoptosis-­
sensitive toward pharmacological apoptosis
inducers upon inhibition of MDM2 by its antagonist nutlin 3 [173, 174].
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
References
13.
1. Swerdlow SH, Campo E, Harris NL, Jaffe ES, Pileri
SA, Stein H et al (2008) Classification of tumours of
haematopoietic and lymphoid tissues, 4th edn. IARC
Press, Lyon
2. Carbone A, Gloghini A, Gaidano G, Franceschi
S, Capello D, Drexler HG et al (1998) Expression
status of BCL-6 and syndecan-1 identifies distinct
14.
61
histogenetic subtypes of Hodgkin’s disease. Blood
92:2220–2228
Greiner A, Tobollik S, Buettner M, Jungnickel B,
Herrmann K, Kremmer E et al (2005) Differential
expression of activation-induced cytidine deaminase
(AID) in nodular lymphocyte-predominant and classical Hodgkin lymphoma. J Pathol 205:541–547
Hartmann S, Eichenauer DA, Plutschow A, Mottok
A, Bob R, Koch K et al (2013) The prognostic
impact of variant histology in nodular lymphocyte-­
predominant Hodgkin lymphoma: a report from
the German Hodgkin Study Group (GHSG). Blood
122:4246–4252
Hansmann ML, Fellbaum C, Hui PK, Zwingers T
(1988) Correlation of content of B cells and Leu7-­
positive cells with subtype and stage in lymphocyte
predominance type Hodgkin's disease. J Cancer Res
Clin Oncol 114:405–410
Kamel OW, Gelb AB, Shibuya RB, Warnke RA
(1993) Leu 7 (CD57) reactivity distinguishes nodular
lymphocyte predominance Hodgkin's disease from
nodular sclerosing Hodgkin's disease, T-cell-­
rich
B-cell lymphoma and follicular lymphoma. Am J
Pathol 142:541–546
Nam-Cha SH, Roncador G, Sanchez-Verde L,
Montes-Moreno S, Acevedo A, Dominguez-Franjo
P et al (2008) PD-1, a follicular T-cell marker useful for recognizing nodular lymphocyte-­predominant
Hodgkin lymphoma. Am J Surg Pathol 32:1252–1257
Liu Y, Abdul Razak FR, Terpstra M, Chan FC, Saber
A, Nijland M et al (2014) The mutational landscape of
Hodgkin lymphoma cell lines determined by wholeexome sequencing. Leukemia 28:2248–2251
Reichel J, Chadburn A, Rubinstein PG, Giulino-­
Roth L, Tam W, Liu Y et al (2015) Flow-sorting
and exome sequencing reveals the oncogenome of
primary Hodgkin and Reed-Sternberg cells. Blood
125:1061–1072
Tiacci E, Ladewig E, Schiavoni G, Penson A, Fortini
E, Pettirossi V et al (2018) Pervasive mutations of
JAK-STAT pathway genes in classical Hodgkin lymphoma. Blood 131:2454–2465
Foss HD, Reusch R, Demel G, Lenz G,
Anagnostopoulos I, Hummel M et al (1999) Frequent
expression of the B-cell-specific activator protein in
Reed-Sternberg cells of classical Hodgkin’s disease
provides further evidence for its B-cell origin. Blood
94:3108–3113
Korkolopoulou P, Cordell J, Jones M, Kaklamanis L,
Tsenga A, Gatter KC et al (1994) The expression of
the B-cell marker mb-1 (CD79a) in Hodgkin’s disease. Histopathology 24:511–515
Kuzu I, Delsol G, Jones M, Gatter KC, Mason DY
(1993) Expression of the Ig-associated heterodimer
(mb-1 and B29) in Hodgkin’s disease. Histopathology
22:141–144
Müschen M, Rajewsky K, Bräuninger A, Baur AS,
Oudejans JJ, Roers A et al (2000) Rare occurrence of
classical Hodgkin’s disease as a T cell lymphoma. J
Exp Med 191:387–394
62
15. Seitz V, Hummel M, Marafioti T, Anagnostopoulos
I, Assaf C, Stein H (2000) Detection of clonal T-cell
receptor gamma-chain gene rearrangements in ReedSternberg cells of classic Hodgkin disease. Blood
95:3020–3024
16. Mani H, Jaffe ES (2009) Hodgkin lymphoma: an
update on its biology with new insights into classification. Clin Lymphoma Myeloma 9:206–216
17. Traverse-Glehen A, Pittaluga S, Gaulard P, Sorbara
L, Alonso MA, Raffeld M et al (2005) Mediastinal
gray zone lymphoma: the missing link between classic Hodgkin’s lymphoma and mediastinal large B-cell
lymphoma. Am J Surg Pathol 29:1411–1421
18. Eckerle S, Brune V, Döring C, Tiacci E, Bohle V,
Sundstrom C et al (2009) Gene expression profiling
of isolated tumour cells from anaplastic large cell
lymphomas: insights into its cellular origin, pathogenesis and relation to Hodgkin lymphoma. Leukemia
23(11):2129–2138
19. Asano N, Yamamoto K, Tamaru J, Oyama T, Ishida
F, Ohshima K et al (2009) Age-related Epstein-Barr
virus (EBV)-associated B-cell lymphoproliferative disorders: comparison with EBV-positive classic Hodgkin lymphoma in elderly patients. Blood
113:2629–2636
20. Kanzler H, Küppers R, Hansmann ML, Rajewsky
K (1996) Hodgkin and Reed-Sternberg cells in
Hodgkin’s disease represent the outgrowth of a dominant tumor clone derived from (crippled) germinal
center B cells. J Exp Med 184:1495–1505
21. Küppers R, Rajewsky K, Zhao M, Simons G,
Laumann R, Fischer R et al (1994) Hodgkin disease:
Hodgkin and Reed-Sternberg cells picked from histological sections show clonal immunoglobulin gene
rearrangements and appear to be derived from B cells
at various stages of development. Proc Natl Acad Sci
U S A 91:10962–10966
22. Marafioti T, Hummel M, Foss H-D, Laumen H,
Korbjuhn P, Anagnostopoulos I et al (2000) Hodgkin
and Reed-Sternberg cells represent an expansion of a
single clone originating from a germinal center B-cell
with functional immunoglobulin gene rearrangements
but defective immunoglobulin transcription. Blood
95:1443–1450
23. Müschen M, Küppers R, Spieker T, Bräuninger A,
Rajewsky K, Hansmann ML (2001) Molecular single-­
cell analysis of Hodgkin- and Reed-Sternberg cells
harboring unmutated immunoglobulin variable region
genes. Lab Invest 81:289–295
24. Küppers R, Zhao M, Hansmann ML, Rajewsky K
(1993) Tracing B cell development in human germinal
centres by molecular analysis of single cells picked
from histological sections. EMBO J 12:4955–4967
25. Küppers R, Rajewsky K (1998) The origin of Hodgkin
and Reed/Sternberg cells in Hodgkin’s disease. Annu
Rev Immunol 16:471–493
26. Lebecque S, de Bouteiller O, Arpin C, Banchereau
J, Liu YJ (1997) Germinal center founder cells display propensity for apoptosis before onset of somatic
mutation. J Exp Med 185:563–571
A. Rosenwald and R. Küppers
27. Bräuninger A, Hansmann ML, Strickler JG, Dummer
R, Burg G, Rajewsky K et al (1999) Identification
of common germinal-center B-cell precursors in
two patients with both Hodgkin’s disease and Non-­
Hodgkin’s lymphoma. N Engl J Med 340:1239–1247
28. Küppers R, Sousa AB, Baur AS, Strickler JG,
Rajewsky K, Hansmann ML (2001) Common
germinal-­center B-cell origin of the malignant cells
in two composite lymphomas, involving classical
Hodgkin's disease and either follicular lymphoma or
B-CLL. Mol Med 7:285–292
29. Marafioti T, Hummel M, Anagnostopoulos I, Foss HD,
Huhn D, Stein H (1999) Classical Hodgkin's disease
and follicular lymphoma originating from the same
germinal center B cell. J Clin Oncol 17:3804–3809
30. Küppers R, Dührsen U, Hansmann ML (2014)
Pathogenesis, diagnosis, and treatment of composite
lymphomas. Lancet Oncol 15:e435–e446
31. Weniger MA, Tiacci E, Schneider S, Arnolds J,
Rüschenbaum S, Duppach J et al (2018) Human
CD30+ B cells represent a unique subset related to
Hodgkin lymphoma cells. J Clin Invest 128:2996–3007
32. Montes-Moreno S, Roncador G, Maestre L, Martinez
N, Sanchez-Verde L, Camacho FI et al (2008) Gcet1
(centerin), a highly restricted marker for a subset of germinal center-derived lymphomas. Blood
111:351–358
33. Natkunam Y, Lossos IS, Taidi B, Zhao S, Lu X, Ding
F et al (2005) Expression of the human germinal
center-associated lymphoma (HGAL) protein, a new
marker of germinal center B-cell derivation. Blood
105:3979–3986
34. Braeuninger A, Küppers R, Strickler JG, Wacker
HH, Rajewsky K, Hansmann ML (1997) Hodgkin
and Reed-Sternberg cells in lymphocyte predominant
Hodgkin disease represent clonal populations of germinal center-derived tumor B cells. Proc Natl Acad
Sci U S A 94:9337–9342
35. Marafioti T, Hummel M, Anagnostopoulos I, Foss
HD, Falini B, Delsol G et al (1997) Origin of nodular
lymphocyte-predominant Hodgkin's disease from a
clonal expansion of highly mutated germinal-center B
cells. N Engl J Med 337:453–458
36. Ohno T, Stribley JA, Wu G, Hinrichs SH, Weisenburger
DD, Chan WC (1997) Clonality in nodular lymphocyte-predominant Hodgkin’s disease. N Engl J Med
337:459–465
37. Brune V, Tiacci E, Pfeil I, Döring C, Eckerle S, van
Noesel CJM et al (2008) Origin and pathogenesis
of nodular lymphocyte-predominant Hodgkin lymphoma as revealed by global gene expression analysis. J Exp Med 205:2251–2268
38. Küppers R, Bräuninger A, Müschen M, Distler V,
Hansmann ML, Rajewsky K (2001) Evidence that
Hodgkin and Reed-Sternberg cells in Hodgkin disease
do not represent cell fusions. Blood 97:818–821
39. Drexler HG, Gignac SM, Hoffbrand AV, Minowada
J (1989) Formation of multinucleated cells in a
Hodgkin’s-disease-derived cell line. Int J Cancer
43:1083–1090
3
Pathology and Molecular Pathology of Hodgkin Lymphoma
40. Newcom SR, Kadin ME, Phillips C (1988) L-428
Reed-Sternberg cells and mononuclear Hodgkin's
cells arise from a single cloned mononuclear cell. Int
J Cell Cloning 6:417–431
41. Ikeda J, Mamat S, Tian T, Wang Y, Rahadiani N,
Aozasa K et al (2010) Tumorigenic potential of
mononucleated small cells of Hodgkin lymphoma cell
lines. Am J Pathol 177:3081–3088
42. Rengstl B, Newrzela S, Heinrich T, Weiser C,
Thalheimer FB, Schmid F et al (2013) Incomplete
cytokinesis and re-fusion of small mononucleated Hodgkin cells lead to giant multinucleated
Reed-Sternberg cells. Proc Natl Acad Sci U S A
110:20729–20734
43. Xavier de Carvalho A, Maiato H, Maia AF, Ribeiro
SA, Pontes P, Bickmore W et al (2015) Reed-­
Sternberg cells form by abscission failure in the
presence of functional Aurora B kinase. PLoS One
10:e0124629
44. Nakashima M, Ishii Y, Watanabe M, Togano T,
Umezawa K, Higashihara M et al (2010) The side
population, as a precursor of Hodgkin and Reed-­
Sternberg cells and a target for nuclear factor-­kappaB
inhibitors in Hodgkin’s lymphoma. Cancer Sci
101:2490–2496
45. Shafer JA, Cruz CR, Leen AM, Ku S, Lu A, Rousseau
A et al (2010) Antigen-specific cytotoxic T lymphocytes can target chemoresistant side-population
tumor cells in Hodgkin lymphoma. Leuk Lymphoma
51:870–880
46. Jansen MP, Hopman AH, Bot FJ, Haesevoets
A, Stevens-Kroef MJ, Arends JW et al (1999)
Morphologically
normal,
CD30-negative
B-lymphocytes with chromosome aberrations in classical Hodgkin’s disease: the progenitor cell of the
malignant clone? J Pathol 189:527–532
47. Spieker T, Kurth J, Küppers R, Rajewsky K,
Bräuninger A, Hansmann ML (2000) Molecular
single-­cell analysis of the clonal relationship of small
Epstein-Barr virus-infected cells and Epstein-­
Barr
virus-harboring Hodgkin and Reed/Sternberg cells in
Hodgkin disease. Blood 96:3133–3138
48. Jones RJ, Gocke CD, Kasamon YL, Miller CB,
Perkins B, Barber JP et al (2009) Circulating clonotypic B cells in classic Hodgkin lymphoma. Blood
113:5920–5926
49. Küppers R (2009) Clonogenic B cells in classic
Hodgkin lymphoma. Blood 114:3970–3971
50. Vockerodt M, Soares M, Kanzler H, Küppers R, Kube
D, Hansmann ML et al (1998) Detection of clonal
Hodgkin and Reed-Sternberg cells with identical
somatically mutated and rearranged VH genes in different biopsies in relapsed Hodgkin’s disease. Blood
92:2899–2907
51. Weber-Matthiesen K, Deerberg J, Poetsch M, Grote
W, Schlegelberger B (1995) Numerical chromosome
aberrations are present within the CD30+ Hodgkin
and Reed-Sternberg cells in 100% of analyzed cases
of Hodgkin’s disease. Blood 86:1464–1468
63
52. Martin-Subero JI, Klapper W, Sotnikova A, CalletBauchu E, Harder L, Bastard C et al (2006)
Chromosomal breakpoints affecting immunoglobulin loci are recurrent in Hodgkin and Reed-Sternberg
cells of classical Hodgkin lymphoma. Cancer Res
66:10332–10338
53. Szymanowska N, Klapper W, Gesk S, Küppers R,
Martin-Subero JI, Siebert R (2008) BCL2 and BCL3
are recurrent translocation partners of the IGH locus.
Cancer Genet Cytogenet 186:110–114
54. Gravel S, Delsol G, Al Saati T (1998) Single-cell
analysis of the t(14;18)(q32;p21) chromosomal
translocation in Hodgkin's disease demonstrates
the absence of this transformation in neoplastic Hodgkin and Reed-Sternberg cells. Blood 91:
2866–2874
55. Poppema S, Kaleta J, Hepperle B (1992) Chromosomal
abnormalities in patients with Hodgkin's disease:
evidence for frequent involvement of the 14q chromosomal region but infrequent bcl-2 gene rearrangement in Reed-Sternberg cells. J Natl Cancer Inst
84:1789–1793
56. Steidl C, Shah SP, Woolcock BW, Rui L, Kawahara
M, Farinha P et al (2011) MHC class II transactivator
CIITA is a recurrent gene fusion partner in lymphoid
cancers. Nature 471:377–381
57. Renné C, Martin-Subero JI, Hansmann ML, Siebert
R (2005) Molecular cytogenetic analyses of immunoglobulin loci in nodular lymphocyte predominant
Hodgkin’s lymphoma reveal a recurrent IGH-BCL6
juxtaposition. J Mol Diagn 7:352–356
58. Wlodarska I, Nooyen P, Maes B, Martin-Subero JI,
Siebert R, Pauwels P et al (2003) Frequent occurrence
of BCL6 rearrangements in nodular lymphocyte predominance Hodgkin lymphoma but not in classical
Hodgkin lymphoma. Blood 101:706–710
59. Wlodarska I, Stul M, De Wolf-Peeters C, Hagemeijer
A (2004) Heterogeneity of BCL6 rearrangements in
nodular lymphocyte predominant Hodgkin’s lymphoma. Haematologica 89:965–972
60. Maggio EM, van den Berg A, de Jong D, Diepstra A,
Poppema S (2003) Low frequency of FAS mutations
in Reed-Sternberg cells of Hodgkin’s lymphoma. Am
J Pathol 162:29–35
61. Müschen M, Re D, Bräuninger A, Wolf J, Hansmann
ML, Diehl V et al (2000) Somatic mutations of the
CD95 gene in Hodgkin and Reed-Sternberg cells.
Cancer Res 60:5640–5643
62. Thomas RK, Schmitz R, Harttrampf AC, Abdil-Hadi
A, Wickenhauser C, Distler V et al (2005) Apoptosis-­
resistant phenotype of classical Hodgkin’s lymphoma
is not mediated by somatic mutations within genes
encoding members of the death-inducing signaling
complex (DISC). Leukemia 19:1079–1082
63. Bose S, Starczynski J, Chukwuma M, Baumforth
K, Wei W, Morgan S et al (2007) Down-regulation
of ATM protein in HRS cells of nodular sclerosis
Hodgkin’s lymphoma in children occurs in the absence
of ATM gene inactivation. J Pathol 213:329–336
64
64. Lespinet V, Terraz F, Recher C, Campo E, Hall J,
Delsol G et al (2005) Single-cell analysis of loss of
heterozygosity at the ATM gene locus in Hodgkin and
Reed-Sternberg cells of Hodgkin’s lymphoma: ATM
loss of heterozygosity is a rare event. Int J Cancer
114:909–916
65. Schmitz R, Thomas RK, Harttrampf AC,
Wickenhauser C, Schultze JL, Hansmann ML et al
(2006) The major subtypes of human B-cell lymphomas lack mutations in BCL-2 family member
BAD. Int J Cancer 119:1738–1740
66. Maggio EM, Stekelenburg E, Van den Berg A,
Poppema S (2001) TP53 gene mutations in Hodgkin
lymphoma are infrequent and not associated with
absence of Epstein-Barr virus. Int J Cancer 94:60–66
67. Montesinos-Rongen M, Roers A, Küppers R,
Rajewsky K, Hansmann M-L (1999) Mutation of
the p53 gene is not a typical feature of Hodgkin and
Reed-Sternberg cells in Hodgkin’s disease. Blood
94:1755–1760
68. Feuerborn A, Moritz C, Von Bonin F, Dobbelstein
M, Trümper L, Sturzenhofecker B et al (2006)
Dysfunctional p53 deletion mutants in cell lines
derived from Hodgkin’s lymphoma. Leuk Lymphoma
47:1932–1940
69. Küpper M, Joos S, Von Bonin F, Daus H, Pfreundschuh
M, Lichter P et al (2001) MDM2 gene amplification
and lack of p53 point mutations in Hodgkin and ReedSternberg cells: results from single-cell polymerase
chain reaction and molecular cytogenetic studies. Br J
Haematol 112:768–775
70. Jardin F, Pujals A, Pelletier L, Bohers E, Camus V,
Mareschal S et al (2016) Recurrent mutations of the
exportin 1 gene (XPO1) and their impact on selective
inhibitor of nuclear export compounds sensitivity in
primary mediastinal B-cell lymphoma. Am J Hematol
91:923–930
71. Abdul Razak FR, Diepstra A, Visser L, van den Berg
A (2016) CD58 mutations are common in Hodgkin
lymphoma cell lines and loss of CD58 expression in
tumor cells occurs in Hodgkin lymphoma patients
who relapse. Genes Immun 17:363–366
72. Schneider M, Schneider S, Zühlke-Jenisch R, Klapper
W, Sundström C, Hartmann S et al (2015) Alterations
of the CD58 gene in classical Hodgkin lymphoma.
Genes Chromosomes Cancer 54:638–645
73. Cabannes E, Khan G, Aillet F, Jarrett RF, Hay RT
(1999) Mutations in the IκΒα gene in Hodgkin’s
disease suggest a tumour suppressor role for IκΒα.
Oncogene 18:3063–3070
74. Emmerich F, Meiser M, Hummel M, Demel G, Foss
HD, Jundt F et al (1999) Overexpression of I kappa
B alpha without inhibition of NF-kappaB activity
and mutations in the I kappa B alpha gene in Reed-­
Sternberg cells. Blood 94:3129–3134
75. Jungnickel B, Staratschek-Jox A, Bräuninger A,
Spieker T, Wolf J, Diehl V et al (2000) Clonal deleterious mutations in the IkBa gene in the malignant cells
in Hodgkin’s disease. J Exp Med 191:395–401
A. Rosenwald and R. Küppers
76. Lake A, Shield LA, Cordano P, Chui DT, Osborne J,
Crae S et al (2009) Mutations of NFKBIA, encoding IkappaBalpha, are a recurrent finding in classical Hodgkin lymphoma but are not a unifying
feature of non-EBV-associated cases. Int J Cancer
125:1334–1342
77. Emmerich F, Theurich S, Hummel M, Haeffker A,
Vry MS, Döhner K et al (2003) Inactivating I kappa B
epsilon mutations in Hodgkin/Reed-Sternberg cells. J
Pathol 201:413–420
78. Mansouri L, Noerenberg D, Young E, Mylonas E,
Abdulla M, Frick M et al (2016) Frequent NFKBIE
deletions are associated with poor outcome in primary
mediastinal B-cell lymphoma. Blood 128:2666–2670
79. Otto C, Giefing M, Massow A, Vater I, Gesk S,
Schlesner M et al (2012) Genetic lesions of the
TRAF3 and MAP3K14 genes in classical Hodgkin
lymphoma. Br J Haematol 157:702–708
80. Schmidt A, Schmitz R, Giefing M, Martin-Subero
JI, Gesk S, Vater I et al (2010) Rare occurrence of
biallelic CYLD gene mutations in classical Hodgkin
lymphoma. Genes Chromosomes Cancer 49:
803–809
81. Joos S, Granzow M, Holtgreve-Grez H, Siebert R,
Harder L, Martin-Subero JI et al (2003) Hodgkin's
lymphoma cell lines are characterized by frequent
aberrations on chromosomes 2p and 9p including
REL and JAK2. Int J Cancer 103:489–495
82. Joos S, Menz CK, Wrobel G, Siebert R, Gesk S, Ohl
S et al (2002) Classical Hodgkin lymphoma is characterized by recurrent copy number gains of the short
arm of chromosome 2. Blood 99:1381–1387
83. Martin-Subero JI, Gesk S, Harder L, Sonoki T,
Tucker PW, Schlegelberger B et al (2002) Recurrent
involvement of the REL and BCL11A loci in classical
Hodgkin lymphoma. Blood 99:1474–1477
84. Barth TF, Martin-Subero JI, Joos S, Menz CK, Hasel
C, Mechtersheimer G et al (2003) Gains of 2p involving the REL locus correlate with nuclear c-Rel protein
accumulation in neoplastic cells of classical Hodgkin
lymphoma. Blood 101:3681–3686
85. Steidl C, Telenius A, Shah SP, Farinha P, Barclay L,
Boyle M et al (2010) Genome-wide copy number
analysis of Hodgkin Reed-Sternberg cells identifies
recurrent imbalances with correlations to treatment
outcome. Blood 116:418–427
86. Martin-Subero JI, Wlodarska I, Bastard C, Picquenot
JM, Höppner J, Giefing M et al (2006) Chromosomal
rearrangements involving the BCL3 locus are recurrent in classical Hodgkin and peripheral T-cell lymphoma. Blood 108:401–402
87. Mathas S, Jöhrens K, Joos S, Lietz A, Hummel F, Janz
M et al (2005) Elevated NF-kappaB p50 complex formation and Bcl-3 expression in classical Hodgkin,
anaplastic large-cell, and other peripheral T-cell lymphomas. Blood 106:4287–4293
88. Kato M, Sanada M, Kato I, Sato Y, Takita J, Takeuchi
K et al (2009) Frequent inactivation of A20 in B-cell
lymphomas. Nature 459:712–716
3
Pathology and Molecular Pathology of Hodgkin Lymphoma
89. Schmitz R, Hartmann S, Giefing M, Mechtersheimer
G, Zuhlke-Jenisch R, Martin-Subero JI et al
(2007) Inactivating mutations of TNFAIP3 (A20)
indicate a tumor suppressor role for A20 in
Hodgkin's ­
lymphoma and primary mediastinal
B cell lymphoma. Haeamtologica. Hematol J 92
(Suppl. 5):41
90. Mosialos G, Birkenbach M, Yalamanchili R,
VanArsdale T, Ware C, Kieff E (1995) The Epstein-­
Barr virus transforming protein LMP1 engages signaling proteins for the tumor necrosis factor receptor
family. Cell 80:389–399
91. Uchida J, Yasui T, Takaoka-Shichijo Y, Muraoka M,
Kulwichit W, Raab-Traub N et al (1999) Mimicry
of CD40 signals by Epstein-Barr virus LMP1 in B
lymphocyte responses. Science 286:300–303
92. Schumacher MA, Schmitz R, Brune V, Tiacci E,
Döring C, Hansmann ML et al (2010) Mutations in
the genes coding for the NF-kappaB regulating factors IkappaBalpha and A20 are uncommon in nodular lymphocyte-predominant Hodgkin’s lymphoma.
Haematologica 95:153–157
93. Mottok A, Renné C, Willenbrock K, Hansmann ML,
Bräuninger A (2007) Somatic hypermutation of
SOCS1 in lymphocyte-predominant Hodgkin lymphoma is accompanied by high JAK2 expression and
activation of STAT6. Blood 110:3387–3390
94. Weniger MA, Melzner I, Menz CK, Wegener S,
Bucur AJ, Dorsch K et al (2006) Mutations of
the tumor suppressor gene SOCS-1 in classical
Hodgkin lymphoma are frequent and associated with
nuclear phospho-STAT5 accumulation. Oncogene
25:2679–2684
95. Gunawardana J, Chan FC, Telenius A, Woolcock
B, Kridel R, Tan KL et al (2014) Recurrent somatic
mutations of PTPN1 in primary mediastinal B cell
lymphoma and Hodgkin lymphoma. Nat Genet
46:329–335
96. Joos S, Küpper M, Ohl S, von Bonin F,
Mechtersheimer G, Bentz M et al (2000) Genomic
imbalances including amplification of the tyrosine
kinase gene JAK2 in CD30+ Hodgkin cells. Cancer
Res 60:549–552
97. Green MR, Monti S, Rodig SJ, Juszczynski P, Currie
T, O'Donnell E et al (2010) Integrative analysis
reveals selective 9p24.1 amplification, increased
PD-1 ligand expression, and further induction via
JAK2 in nodular sclerosing Hodgkin lymphoma and
primary mediastinal large B-cell lymphoma. Blood
116:3268–3277
98. Rui L, Emre NC, Kruhlak MJ, Chung HJ, Steidl
C, Slack G et al (2010) Cooperative epigenetic
modulation by cancer amplicon genes. Cancer Cell
18:590–605
99. Van Roosbroeck K, Cox L, Tousseyn T, Lahortiga I,
Gielen O, Cauwelier B et al (2011) JAK2 rearrangements, including the novel SEC31A-JAK2 fusion,
are recurrent in classical Hodgkin lymphoma. Blood
117:4056–4064
65
100. Hartmann S, Martin-Subero JI, Gesk S, Husken J,
Giefing M, Nagel I et al (2008) Detection of genomic
imbalances in microdissected Hodgkin and Reed-­
Sternberg cells of classical Hodgkin's lymphoma
by array-based comparative genomic hybridization.
Haematologica 93:1318–1326
101. Hartmann S, Schuhmacher B, Rausch T, Fuller L,
Döring C, Weniger M et al (2016) Highly recurrent
mutations of SGK1, DUSP2 and JUNB in nodular lymphocyte predominant Hodgkin lymphoma.
Leukemia 30:844–853
102. Re D, Müschen M, Ahmadi T, Wickenhauser C,
Staratschek-Jox A, Holtick U et al (2001) Oct-2 and
Bob-1 deficiency in Hodgkin and Reed Sternberg
cells. Cancer Res 61:2080–2084
103. Stein H, Marafioti T, Foss HD, Laumen H, Hummel
M, Anagnostopoulos I et al (2001) Down-regulation
of BOB.1/OBF.1 and Oct2 in classical Hodgkin disease but not in lymphocyte predominant Hodgkin
disease correlates with immunoglobulin transcription. Blood 97:496–501
104. Watanabe K, Yamashita Y, Nakayama A, Hasegawa
Y, Kojima H, Nagasawa T et al (2000) Varied B-cell
immunophenotypes of Hodgkin/Reed-Sternberg
cells in classic Hodgkin’s disease. Histopathology
36:353–361
105. Schwering I, Bräuninger A, Klein U, Jungnickel
B, Tinguely M, Diehl V et al (2003) Loss of the
B-lineage-specific gene expression program in
Hodgkin and Reed-Sternberg cells of Hodgkin lymphoma. Blood 101:1505–1512
106. Carbone A, Gloghini A, Larocca LM, Antinori
A, Falini B, Tirelli U et al (1999) Human immunodeficiency virus-associated Hodgkin’s disease
derives from post-germinal center B cells. Blood
93:2319–2326
107. Tiacci E, Döring C, Brune V, van Noesel CJ, Klapper
W, Mechtersheimer G et al (2012) Analyzing primary Hodgkin and Reed-Sternberg cells to capture
the molecular and cellular pathogenesis of classical
Hodgkin lymphoma. Blood 120:4609–4620
108. Poppema S (1996) Immunology of Hodgkin’s disease. Baillieres Clin Haematol 9:447–457
109. Carbone A, Gloghini A, Gruss HJ, Pinto A (1995)
CD40 ligand is constitutively expressed in a subset of T cell lymphomas and on the microenvironmental reactive T cells of follicular lymphomas and
Hodgkin’s disease. Am J Pathol 147:912–922
110. Torlakovic E, Tierens A, Dang HD, Delabie J (2001)
The transcription factor PU.1, necessary for B-cell
development is expressed in lymphocyte predominance, but not classical Hodgkin’s disease. Am J
Pathol 159:1807–1814
111. Bohle V, Döring C, Hansmann ML, Küppers R
(2013) Role of early B-cell factor 1 (EBF1) in
Hodgkin lymphoma. Leukemia 27:671–679
112. Overbeck BM, Martin-Subero JI, Ammerpohl O,
Klapper W, Siebert R, Giefing M (2012) ETS1
encoding a transcription factor involved in B-cell
A. Rosenwald and R. Küppers
66
113.
114.
115.
116.
117.
118.
119.
120.
121.
122.
123.
124.
differentiation is recurrently deleted and downregulated in classical Hodgkin’s lymphoma.
Haematologica 97:1612–1614
Küppers R, Klein U, Schwering I, Distler V,
Bräuninger A, Cattoretti G et al (2003) Identification
of Hodgkin and Reed-Sternberg cell-specific
genes by gene expression profiling. J Clin Invest
111:529–537
Mathas S, Janz M, Hummel F, Hummel M, Wollert-­
Wulf B, Lusatis S et al (2006) Intrinsic inhibition of
transcription factor E2A by HLH proteins ABF-1 and
Id2 mediates reprogramming of neoplastic B cells in
Hodgkin lymphoma. Nat Immunol 7:207–215
Renné C, Martin-Subero JI, Eickernjager M,
Hansmann ML, Küppers R, Siebert R et al (2006)
Aberrant expression of ID2, a suppressor of B-cell-­
specific gene expression, in Hodgkin’s lymphoma.
Am J Pathol 169:655–664
Hacker C, Kirsch RD, Ju XS, Hieronymus T, Gust
TC, Kuhl C et al (2003) Transcriptional profiling
identifies Id2 function in dendritic cell development.
Nat Immunol 4:380–386
Yokota Y, Mansouri A, Mori S, Sugawara S, Adachi
S, Nishikawa S et al (1999) Development of
peripheral lymphoid organs and natural killer cells
depends on the helix-loop-helix inhibitor Id2. Nature
397:702–706
Jundt F, Acikgoz O, Kwon SH, Schwarzer R,
Anagnostopoulos I, Wiesner B et al (2008) Aberrant
expression of Notch1 interferes with the B-lymphoid
phenotype of neoplastic B cells in classical Hodgkin
lymphoma. Leukemia 22:1587–1594
Jundt F, Anagnostopoulos I, Förster R, Mathas S,
Stein H, Dörken B (2002) Activated Notch 1 signaling promotes tumor cell proliferation and survival in
Hodgkin and anaplastic large cell lymphoma. Blood
99:3398–3403
Köchert K, Ullrich K, Kreher S, Aster JC, Kitagawa
M, Johrens K et al (2011) High-level expression of
Mastermind-like 2 contributes to aberrant activation
of the NOTCH signaling pathway in human lymphomas. Oncogene 30(15):1831–1840
Scheeren FA, Diehl SA, Smit LA, Beaumont T,
Naspetti M, Bende RJ et al (2008) IL-21 is expressed
in Hodgkin lymphoma and activates STAT5; evidence that activated STAT5 is required for Hodgkin
lymphomagenesis. Blood 111:4706–4715
Stanelle J, Döring C, Hansmann ML, Küppers R
(2010) Mechanisms of aberrant GATA3 expression in classical Hodgkin lymphoma and its consequences for the cytokine profile of Hodgkin and
Reed/Sternberg cells. Blood 116:4202–4211
Doerr JR, Malone CS, Fike FM, Gordon MS,
Soghomonian SV, Thomas RK et al (2005) Patterned
CpG methylation of silenced B cell gene promoters in classical Hodgkin lymphoma-derived and
primary effusion lymphoma cell lines. J Mol Biol
350:631–640
Ushmorov A, Leithäuser F, Sakk O, Weinhausel
A, Popov SW, Möller P et al (2006) Epigenetic
125.
126.
127.
128.
129.
130.
131.
132.
133.
134.
135.
136.
137.
processes play a major role in B-cell-specific gene
silencing in classical Hodgkin lymphoma. Blood
107:2493–2500
Ammerpohl O, Haake A, Pellissery S, Giefing M,
Richter J, Balint B et al (2012) Array-based DNA
methylation analysis in classical Hodgkin lymphoma
reveals new insights into the mechanisms underlying silencing of B cell-specific genes. Leukemia
26:185–188
Dukers DF, van Galen JC, Giroth C, Jansen P,
Sewalt RG, Otte AP et al (2004) Unique polycomb
gene expression pattern in Hodgkin's lymphoma
and Hodgkin’s lymphoma-derived cell lines. Am J
Pathol 164:873–881
Raaphorst FM, van Kemenade FJ, Blokzijl T, Fieret
E, Hamer KM, Satijn DP et al (2000) Coexpression
of BMI-1 and EZH2 polycomb group genes in Reed-­
Sternberg cells of Hodgkin’s disease. Am J Pathol
157:709–715
Sanchez-Beato M, Sanchez E, Garcia JF, Perez-­
Rosado A, Montoya MC, Fraga M et al (2004)
Abnormal PcG protein expression in Hodgkin’s
lymphoma. Relation with E2F6 and NFkappaB transcription factors. J Pathol 204:528–537
Schneider EM, Torlakovic E, Stuhler A, Diehl V,
Tesch H, Giebel B (2004) The early transcription
factor GATA-2 is expressed in classical Hodgkin’s
lymphoma. J Pathol 204:538–545
Lamprecht B, Walter K, Kreher S, Kumar R,
Hummel M, Lenze D et al (2010) Derepression of
an endogenous long terminal repeat activates the
CSF1R proto-oncogene in human lymphoma. Nat
Med 16:571–579
Yuki H, Ueno S, Tatetsu H, Niiro H, Iino T, Endo S
et al (2013) PU.1 is a potent tumor suppressor in classical Hodgkin lymphoma cells. Blood 121:962–970
Guan H, Xie L, Wirth T, Ushmorov A (2016)
Repression of TCF3/E2A contributes to Hodgkin
lymphomagenesis. Oncotarget 7:36854–36864
Xie L, Ushmorov A, Leithäuser F, Guan H, Steidl
C, Farbinger J et al (2012) FOXO1 is a tumor suppressor in classical Hodgkin lymphoma. Blood
119:3503–3511
Du J, Neuenschwander M, Yu Y, Dabritz
JH, Neuendorff NR, Schleich K et al (2017)
Pharmacological restoration and therapeutic targeting of the B-cell phenotype in classical Hodgkin
lymphoma. Blood 129:71–81
Bargou RC, Emmerich F, Krappmann D, Bommert
K, Mapara MY, Arnold W et al (1997) Constitutive
nuclear factor-kappaB-RelA activation is required
for proliferation and survival of Hodgkin’s disease
tumor cells. J Clin Invest 100:2961–2969
Carbone A, Gloghini A, Gattei V, Aldinucci D,
Degan M, De Paoli P et al (1995) Expression of
functional CD40 antigen on Reed-Sternberg cells
and Hodgkin’s disease cell lines. Blood 85:780–789
Chiu A, Xu W, He B, Dillon SR, Gross JA, Sievers
E et al (2007) Hodgkin lymphoma cells express
TACI and BCMA receptors and generate survival
3
Pathology and Molecular Pathology of Hodgkin Lymphoma
138.
139.
140.
141.
142.
143.
144.
145.
146.
147.
148.
149.
and proliferation signals in response to BAFF and
APRIL. Blood 109:729–739
Fiumara P, Snell V, Li Y, Mukhopadhyay A, Younes
M, Gillenwater AM et al (2001) Functional expression of receptor activator of nuclear factor kappaB in
Hodgkin disease cell lines. Blood 98:2784–2790
Molin D, Fischer M, Xiang Z, Larsson U, Harvima
I, Venge P et al (2001) Mast cells express functional
CD30 ligand and are the predominant CD30L-­
positive cells in Hodgkin’s disease. Br J Haematol
114:616–623
Schwab U, Stein H, Gerdes J, Lemke H, Kirchner H,
Schaadt M et al (1982) Production of a monoclonal
antibody specific for Hodgkin and Sternberg-Reed
cells of Hodgkin’s disease and a subset of normal
lymphoid cells. Nature 299:65–67
Hirsch B, Hummel M, Bentink S, Fouladi F, Spang
R, Zollinger R et al (2008) CD30-induced signaling is absent in Hodgkin’s cells but present in anaplastic large cell lymphoma cells. Am J Pathol 172:
510–520
Horie R, Watanabe T, Morishita Y, Ito K, Ishida T,
Kanegae Y et al (2002) Ligand-independent signaling by overexpressed CD30 drives NF-kappaB activation in Hodgkin-Reed-Sternberg cells. Oncogene
21:2493–2503
Kilger E, Kieser A, Baumann M, Hammerschmidt
W (1998) Epstein-Barr virus-mediated B-cell proliferation is dependent upon latent membrane protein 1, which simulates an activated CD40 receptor.
EMBO J 17:1700–1709
Baus D, Pfitzner E (2006) Specific function of
STAT3, SOCS1, and SOCS3 in the regulation of
proliferation and survival of classical Hodgkin lymphoma cells. Int J Cancer 118:1404–1413
Kube D, Holtick U, Vockerodt M, Ahmadi T,
Behrmann I, Heinrich PC et al (2001) STAT3 is
constitutively activated in Hodgkin cell lines. Blood
98:762–770
Skinnider BF, Elia AJ, Gascoyne RD, Patterson
B, Trümper L, Kapp U et al (2002) Signal transducer and activator of transcription 6 is frequently
activated in Hodgkin and Reed-Sternberg cells of
Hodgkin lymphoma. Blood 99:618–626
Kapp U, Yeh WC, Patterson B, Elia AJ, Kagi D, Ho
A et al (1999) Interleukin 13 is secreted by and stimulates the growth of Hodgkin and Reed-Sternberg
cells. J Exp Med 189:1939–1946
Skinnider BF, Elia AJ, Gascoyne RD, Trumper LH,
von Bonin F, Kapp U et al (2001) Interleukin 13 and
interleukin 13 receptor are frequently expressed by
Hodgkin and Reed-Sternberg cells of Hodgkin lymphoma. Blood 97:250–255
Hinz M, Lemke P, Anagnostopoulos I, Hacker C,
Krappmann D, Mathas S et al (2002) Nuclear factor kappaB-dependent gene expression profiling of
Hodgkin's disease tumor cells, pathogenetic significance, and link to constitutive signal transducer
and activator of transcription 5a activity. J Exp Med
196:605–617
67
150. Lamprecht B, Kreher S, Anagnostopoulos I, Johrens
K, Monteleone G, Jundt F et al (2008) Aberrant
expression of the Th2 cytokine IL-21 in Hodgkin
lymphoma cells regulates STAT3 signaling and
attracts Treg cells via regulation of MIP-3{alpha}.
Blood 112:3339–3347
151. Blume-Jensen P, Hunter T (2001) Oncogenic kinase
signalling. Nature 411:355–365
152. Renné C, Willenbrock K, Küppers R, Hansmann
M-L, Bräuninger A (2005) Autocrine and paracrine activated receptor tyrosine kinases in classical
Hodgkin lymphoma. Blood 105:4051–4059
153. Teofili L, Di Febo AL, Pierconti F, Maggiano N,
Bendandi M, Rutella S et al (2001) Expression of
the c-met proto-oncogene and its ligand, hepatocyte growth factor, in Hodgkin disease. Blood
97:1063–1069
154. Renné C, Willenbrock K, Martin-Subero JI, Hinsch
N, Döring C, Tiacci E et al (2007) High expression of
several tyrosine kinases and activation of the PI3K/
AKT pathway in mediastinal large B cell lymphoma
reveals further similarities to Hodgkin lymphoma.
Leukemia 21:780–787
155. Renné C, Hinsch N, Willenbrock K, Fuchs M,
Klapper W, Engert A et al (2007) The aberrant
coexpression of several receptor tyrosine kinases is
largely restricted to EBV-negative cases of classical
Hodgkin’s lymphoma. Int J Cancer 120:2504–2509
156. Renne C, Minner S, Küppers R, Hansmann ML,
Bräuninger A (2008) Autocrine NGFbeta/TRKA
signalling is an important survival factor for Hodgkin
lymphoma derived cell lines. Leuk Res 32:163–167
157. Nagel S, Burek C, Venturini L, Scherr M, Quentmeier
H, Meyer C et al (2007) Comprehensive analysis of homeobox genes in Hodgkin lymphoma cell
lines identifies dysregulated expression of HOXB9
mediated via ERK5 signaling and BMI1. Blood
109:3015–3023
158. Zheng B, Fiumara P, Li YV, Georgakis G, Snell
V, Younes M et al (2003) MEK/ERK pathway is
aberrantly active in Hodgkin disease: a signaling pathway shared by CD30, CD40, and RANK
that regulates cell proliferation and survival. Blood
102:1019–1027
159. Mathas S, Hinz M, Anagnostopoulos I, Krappmann
D, Lietz A, Jundt F et al (2002) Aberrantly expressed
c-Jun and JunB are a hallmark of Hodgkin lymphoma cells, stimulate proliferation and synergize
with NF-kappa B. EMBO J 21:4104–4113
160. Juszczynski P, Ouyang J, Monti S, Rodig SJ,
Takeyama K, Abramson J et al (2007) The
AP1-dependent secretion of galectin-1 by Reed
Sternberg cells fosters immune privilege in classical Hodgkin lymphoma. Proc Natl Acad Sci U S A
104:13134–13139
161. Watanabe M, Ogawa Y, Ito K, Higashihara M, Kadin
ME, Abraham LJ et al (2003) AP-1 mediated relief
of repressive activity of the CD30 promoter microsatellite in Hodgkin and Reed-Sternberg cells. Am J
Pathol 163:633–641
68
162. Lollies A, Hartmann S, Schneider M, Bracht T, Weiss
AL, Arnolds J et al (2018) An oncogenic axis of
STAT-mediated BATF3 upregulation causing MYC
activity in classical Hodgkin lymphoma and anaplastic large cell lymphoma. Leukemia 32:92–101
163. Vrzalikova K, Ibrahim M, Vockerodt M, Perry T,
Margielewska S, Lupino L et al (2018) S1PR1 drives
a feedforward signalling loop to regulate BATF3 and
the transcriptional programme of Hodgkin lymphoma cells. Leukemia 32:214–223
164. Dutton A, Reynolds GM, Dawson CW, Young
LS, Murray PG (2005) Constitutive activation of
phosphatidyl-inositide 3 kinase contributes to the
survival of Hodgkin's lymphoma cells through a
mechanism involving Akt kinase and mTOR. J
Pathol 205:498–506
165. Georgakis GV, Li Y, Rassidakis GZ, Medeiros LJ,
Mills GB, Younes A (2006) Inhibition of the phosphatidylinositol-­
3 kinase/Akt promotes G1 cell
cycle arrest and apoptosis in Hodgkin lymphoma. Br
J Haematol 132:503–511
166. Dutton A, O'Neil JD, Milner AE, Reynolds GM,
Starczynski J, Crocker J et al (2004) Expression
of the cellular FLICE-inhibitory protein (c-FLIP)
protects Hodgkin’s lymphoma cells from autonomous Fas-mediated death. Proc Natl Acad Sci U S A
101:6611–6616
167. Mathas S, Lietz A, Anagnostopoulos I, Hummel
F, Wiesner B, Janz M et al (2004) c-FLIP mediates resistance of Hodgkin/Reed-Sternberg cells
to death receptor-induced apoptosis. J Exp Med
199:1041–1052
A. Rosenwald and R. Küppers
168. Re D, Hofmann A, Wolf J, Diehl V, Staratschek-­
Jox A (2000) Cultivated H-RS cells are resistant to
CD95L-mediated apoptosis despite expression of
wild-type CD95. Exp Hematol 28:31–35
169. Chu WS, Aguilera NS, Wei MQ, Abbondanzo SL
(1999) Antiapoptotic marker Bcl-X(L), expression on Reed-Sternberg cells of Hodgkin’s disease
using a novel monoclonal marker, YTH-2H12. Hum
Pathol 30:1065–1070
170. Kashkar H, Haefs C, Shin H, Hamilton-Dutoit SJ,
Salvesen GS, Krönke M et al (2003) XIAP-mediated
caspase inhibition in Hodgkin’s lymphoma-derived
B cells. J Exp Med 198:341–347
171. Kashkar H, Seeger JM, Hombach A, Deggerich A,
Yazdanpanah B, Utermohlen O et al (2006) XIAP
targeting sensitizes Hodgkin lymphoma cells for
cytolytic T-cell attack. Blood 108:3434–3440
172. Sanchez-Beato M, Piris MA, Martinez-Montero
JC, Garcia JF, Villuendas R, Garcia FJ et al (1996)
MDM2 and p21WAF1/CIP1, wild-type p53-induced
proteins, are regularly expressed by Sternberg-Reed
cells in Hodgkin's disease. J Pathol 180:58–64
173. Drakos E, Thomaides A, Medeiros LJ, Li J,
Leventaki V, Konopleva M et al (2007) Inhibition of
p53-murine double minute 2 interaction by nutlin­3A stabilizes p53 and induces cell cycle arrest and
apoptosis in Hodgkin lymphoma. Clin Cancer Res
13:3380–3387
174. Janz M, Stuhmer T, Vassilev LT, Bargou RC (2007)
Pharmacologic activation of p53-dependent and
p53-independent apoptotic pathways in Hodgkin/
Reed-Sternberg cells. Leukemia 21:772–779
4
Microenvironment, Cross-Talk,
and Immune Escape Mechanisms
Lydia Visser, Johanna Veldman, Sibrand Poppema,
Anke van den Berg, and Arjan Diepstra
Contents
4.1
4.1.1
4.1.2
4.1.3
4.1.4
4.1.5
4.1.6
4.1.7
Microenvironment Hodgkin Lymphoma Subtypes Epstein-Barr Virus Human Immunodeficiency Virus T Cell Subsets in cHL T Cell Subsets in NLPHL Fibrosis and Sclerosis Eosinophils, Plasma Cells, Mast Cells, and B Cells 70
70
70
71
71
73
74
74
4.2
4.2.1
4.2.2
ross-Talk between HRS Cells and Microenvironment C
Factors Supporting Tumor Growth Shaping the Environment 74
74
76
4.3
4.3.1
4.3.2
4.3.3
4.3.4
4.3.5
I mmune Escape Mechanisms Antigen Presentation HLA Class I Expression HLA Class II Expression Immune Checkpoints Immunosuppression 77
77
78
78
79
80
4.4
Prognostic Impact of the Microenvironment 81
4.5
Conclusion 81
References 81
L. Visser · J. Veldman · S. Poppema
A. van den Berg · A. Diepstra (*)
Department of Pathology and Medical Biology,
University Medical Center Groningen, University
of Groningen, Groningen, The Netherlands
e-mail: l.visser@umcg.nl; j.veldman@umcg.nl;
s.poppema@rug.nl; a.van.den.berg01@umcg.nl;
a.diepstra@umcg.nl
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_4
69
L. Visser et al.
70
4.1
Microenvironment
4.1.1
Hodgkin Lymphoma Subtypes
When discussing the microenvironment in Hodgkin
lymphoma (HL), it is important to recognize the
different HL subtypes described by the WHO classification [1, 2]. The classical HL (cHL) subtypes
are defined in large part by the composition of the
reactive infiltrate (Table 4.1). The most prevalent
subtype is the nodular sclerosis type that consists of
a nodular background with thick fibrotic bands,
usually with a thickened lymph node capsule. In
addition to the lacunar type of Hodgkin/ReedSternberg (HRS) cells, there is a microenvironment
consisting of T cells, eosinophils, and histiocytes,
with a variable admixture of neutrophils, plasma
cells, fibroblasts, and mast cells. The second most
common subtype is mixed cellularity, which is
defined by the presence of typical HRS cells and a
diffuse infiltrate of T cells, eosinophils, histiocytes,
and plasma cells, sometimes with the formation of
granuloma-like clusters or granulomas (Fig. 4.1).
Lymphocyte-rich cHL also comprises typical HRS
cells in a nodular or diffuse microenvironment and
small B and/or T lymphocytes dominating the
background, sometimes with admixture of histiocytes. Granulocytes are not present in this subtype.
The rare lymphocyte-depleted subtype harbors a
high percentage of HRS cells in a background consisting of fibroblasts and a low number of T cells.
Nodular lymphocyte predominance (NLP) HL is
an entity that is fundamentally different from
cHL. The morphology may closely resemble that
of the nodular variant of the classical lymphocyterich subtype, both involving follicular areas with
many small B cells, but NLPHL can also show
other growth patterns [3]. The characteristics of the
tumor cells and the T cells in NLPHL are different
from cHL. HRS cells show a loss of B cell phenotype, while in NLPHL the lymphocyte-predominant (LP) tumor cells share many markers with
germinal center B cells. The T cells in cHL have
features of paracortical T cells, while those in
NLPHL are similar to germinal center T cells (T
follicular helper cells) [4–6].
4.1.2
Epstein-Barr Virus
The presence of latent Epstein-Barr virus (EBV)
genomes in HRS cells appears to influence the
composition of the microenvironment. Positive
EBV status is strongly associated with the mixed
Fig. 4.1 The microenvironment in mixed cellularity classical Hodgkin lymphoma. RS classical Reed-Sternberg
cell, H mononuclear Hodgkin tumor cell, T T lymphocyte,
Hi histiocyte, E eosinophil, P plasma cell. Hematoxylin
and eosin staining
Table 4.1 Composition of the microenvironment in different Hodgkin lymphoma (HL) subtypes
Subtype
Nodular sclerosis
EBV (%) Background
10–40
Nodular + fibrosis
T cells
CD4 > CD8, Th2,
Treg > Th1
Mixed cellularity
75
Diffuse
Lymphocyte rich
Lymphocyte depleted
(including HIV+)
Nodular lymphocyte
predominant
40–80
80–100
Nodular or diffuse
Diffuse
CD4 > CD8, Th2,
Treg > Th1
CD4 > CD8
–
~2
Nodular (+diffuse)
Th2, PD1+/CD57+
Tfh, CD4+/8+
Other cells
Eosinophils, histiocytes,
fibroblasts, B cells, mast cells
(neutrophils)
Eosinophils, histiocytes,
plasma cells, B cells
Histiocytes
Fibroblasts
Histiocytes, B cells
4
Microenvironment, Cross-Talk, and Immune Escape Mechanisms
cellularity subtype (~75% EBV-positive) and is
almost always absent in NLPHL [7]. Depending
on the geographic locale, EBV is present in the
HRS cells in 10–40% in nodular sclerosis cases.
The percentage of EBV-positive lymphocyte-rich
cHL cases is not very clear, but is probably
between 40% and 80%. EBV infects more than
90% of the world population and establishes a
lifelong latent infection in B cells in its host.
Potent cytotoxic immune responses keep the
number of EBV-infected B cells at approximately
1/100,000 and usually prevents EBV-driven
malignant transformation in immunocompetent
individuals. Accordingly, EBV-positive cHL
cases contain slightly more CD8+ cytotoxic T
cells in the reactive background compared to
EBV-negative cHL cases [8].
4.1.3
Human Immunodeficiency
Virus
In patients with an impaired immune response,
cHL occurs more frequently. After solid organ
transplantation, there is a small increase in the
incidence of cHL that can largely be attributed to
EBV-positive cHL. Human immunodeficiency
virus (HIV)-infected individuals have an approximate 10 times increased risk of developing cHL
[9]. In comparison to non-HIV-associated cHL,
these tumors are more often EBV-associated,
mixed cellularity and lymphocyte depletion subtypes, and usually contain more tumor cells. This
may reflect a functional defect in the immune
response, in particular to EBV, presumably caused
by the impairment of CD4+ T cells by HIV. On
the other hand, the importance of CD4+ T cells
for supporting the growth of HRS cells is also
illustrated in HIV-positive patients, in which an
increase in the incidence of HIV-associated cHL
has been observed after introduction of highly
active antiretroviral therapy (HAART) [10].
4.1.4
T Cell Subsets in cHL
A unifying feature of the reactive infiltrate in virtually all cHL subtypes is the presence of large
71
numbers of CD4+ T cells. Besides being widely
distributed in the background, these CD4+ T
cells form a tight rosette around the tumor cells.
T cells within these rosettes often have a distinct
phenotype, which is different from the phenotype
of the T cells that are located further away from
the cHL tumor cells (Fig. 4.2).
In general, CD4+ T cells are divided into naive
(CD45RA+) and memory (CD45RO+) subsets
based on whether they have previously been stimulated by an antigen or not. A large subset of CD4+
T cells consists of the so-called helper T (Th) cells;
these cells play an important role in helping other
cells to induce an effective immune response. Th
cells can be further divided into Th0 (naive), Th1
(cellular response), Th2 (humoral response), Th17
(IL-17 producing), and Treg (regulating responses
of other cells) cells. The Treg cells can be further
divided into Th3 (transforming growth factor-β
(TGF-β) producing), Tr1 (IL-10 producing), and
CD4+CD25+ Treg (originating from the thymus)
subpopulations. Some, but not all, Treg cells
express the transcription factor FoxP3.
The T cells in cHL consist mainly of CD4+ T
cells that have a memory phenotype (CD45RO+)
and express several activation markers including
CD28, CD38, CD69, CD71, CD25, and HLA-DR,
as well as markers like CD28, CTLA-4, and
CD40L. However, these T cells lack expression
of CD26 [11]. This lack of CD26 expression is
most striking in the areas surrounding the tumor
cells. CD26, dipeptidyl peptidase IV, regulates
proteolytic processing of several chemokines,
e.g., CCL5 (Rantes), CCL11 (Eotaxin), and
CCL22 (MDC) [12]. CD26 is also associated
with adenosine deaminase (ADA) and CD45RO
and when interacting with anti-CD26 antibodies
leads to enhanced T cell activation through triggering of the T cell receptor [13]. CD26 is preferentially expressed on CD4+CD45RO+ cells and
is normally upregulated after activation. However,
CD26 cannot be upregulated by ex vivo activation of the CD26-negative cells from cHL lesions.
In general, a high CD26 expression level correlates with a Th1 subtype.
The transcription factor expression pattern
indicates that the CD4+ T cells in cHL are predominantly Th2 (c-Maf) and Treg (FoxP3) [4, 14].
72
L. Visser et al.
Fig. 4.2 Shaping the microenvironment in classical
Hodgkin lymphoma (HL). Immunohistochemistry of
classical HL cases. In the upper panel left, strong and
specific staining of Hodgkin/Reed-Sternberg (HRS)
cells for chemokine CCL17 (TARC). This chemokine
attracts CCR4+ lymphocytes (upper panel right). A
large proportion of reactive T cells are Treg cells, as
shown by positive staining for transcription factor
FoxP3 (lower panel left) and activation marker CD25
(lower panel right)
The CD4+CD26− T cell subset in cHL has
reduced mRNA levels of Th1- and Th2-associated
cytokines in comparison to the CD4+CD26+ T
cells from cHL and CD4+ T cells (both CD26−
and CD26+) in reactive lymph nodes [15]. Based
on much higher mRNA expression levels of
IL-2RA (CD25), CCR4, FoxP3, CTLA4,
TNFRSF4 (OX-40), and TNFRSF18 (GITR)
observed in the CD4+CD26− T cells from cHL, it
has been postulated that these cells have a Treg
phenotype (Fig. 4.2). In addition, mildly enhanced
IL-17 levels can be observed both in CD4+CD26−
and CD4+CD26+ T cells from cHL in comparison
to the T cells from tonsil. Upon stimulation, the
CD4+CD26− T cells fail to induce expression of
cytokines, suggesting that the T cell population
rosetting around the HRS cells or located in the
direct vicinity of the HRS cells have an anergic
phenotype (i.e., do not respond to stimulation)
[15]. Immunohistochemistry for several Treg-­
associated molecules demonstrates that the rosetting T cells in cHL express GITR, CCR4, and
CD25, but not FoxP3. Scattered FoxP3-positive
cells are present in the infiltrate, but only rarely in
the direct vicinity of the HRS cells, whereas
CTLA-4 shows a more diffuse presence [15].
Likewise, a small number of scattered IL-17-­
positive cells can be found in the reactive infiltrate.
Th17 cells are generally pro-inflammatory, but
given the abundance of TGF-β and IL-6 in the
Hodgkin microenvironment, the observed IL-17-­
positive cells might be yet another type of regulatory cells, termed Treg17 cells. Although the vast
majority of studies indicate that the CD4+ T cells
in cHL are (anergic) Th2 cells and Treg cells, some
studies showed a predominant Th1-type pattern in
whole lymph node cell suspensions, with a mild
increase in EBV-positive cHL [16]. These findings
4
Microenvironment, Cross-Talk, and Immune Escape Mechanisms
are not contradictory, as these Th1-like cells are
located mainly outside the areas where R-S cells
and T cell rosettes are found [17, 18].
73
The CD4+ T cells in NLPHL resemble the CD4+
T cells in cHL, regarding the expression of
CD45RO, CD69, CTLA4, and CD28 and lack of
CD26. However, the T cells in NLPHL do not
express CD40L, and a significant proportion of
the T cells that immediately surround the LP
cells express CD57 and PD-1 [19, 20]. Similar to
the Th2 cells in cHL, the rosetting cells in
NLPHL strongly express the Th2-associated
transcription factor c-Maf (Fig. 4.3; [4]).
Characterization of the cytokine profile of the
CD4+CD57+ T cell subset shows lack of IL-2
and IL-4 mRNA, but elevated interferon-γ (IFN-­
γ) mRNA levels in comparison to CD57+ T cells
from tonsils. Stimulation of these cells fails to
induce IL-2 and IL-4 mRNA levels [21], which is
similar to the lack of cytokine induction upon
stimulation of the CD26− T cells in cHL. In normal tissue, CD4+CD57+ T cells are found almost
exclusively in the light zone of reactive germinal
centers and also lack CD40L expression. CD57 is
known as an activation marker, but it has also
been demonstrated to be a marker for senescent
cells. Senescence is the phenomenon by which
normal diploid cells lose the ability to divide,
normally after about 50 cell divisions. PD-1+ T
follicular helper cells are present in NLPHL;
Fig. 4.3 T cells in nodular lymphocyte-predominant
Hodgkin lymphoma (NLPHL). Immunohistochemistry of
an NLPHL case showing a variable but usually high number of reactive T cells that express CD57. In this case
these T cells form a rosette around the tumor cells (upper
panel left). The CD57+ T cells also express the transcription factor c-Maf, indicating a Th2-type nature (upper
panel right). In addition, these cells express the T follicular helper cell-associated markers PD-1 (lower panel left)
and BCL-6 (lower panel right)
4.1.5
T Cell Subsets in NLPHL
L. Visser et al.
74
these cells normally provide help to B cells during the germinal center reaction. Another cell
population, consisting of CD4+CD8+ dual
­positive T cells, has been reported to be present in
more than 50% of NLPHL tumors. A similar
population was found in reactive lymph nodes
with progressively transformed germinal centers,
which can also be seen in conjunction with
NLPHL. In normal peripheral blood, they constitute 1–2% of T cells. The function of these cells
is ill defined, and currently they are considered to
be potent immunosuppressors and/or to have
high cytotoxic potential [22].
4.1.6
Fibrosis and Sclerosis
The presence of bands of collagen surrounding
nodules and blood vessels is typical of the nodular sclerosis subtype. Several factors can
induce the activation of fibroblasts and the subsequent deposition of extracellular matrix proteins. The Th2 cells in cHL might provide a
profibrogenic microenvironment by the production of the Th2 cytokine IL-13. IL-13 is
expressed at a higher level in nodular sclerosis
than in mixed cellularity cHL. Moreover, the
percentage of IL-13 receptor-positive fibroblasts is increased in nodular sclerosis cHL
cases [23]. IL-13 stimulates collagen synthesis
in vitro and also stimulates the production of
TGF-β, another potent stimulator of fibrosis.
TGF-β can interact with basic fibroblast growth
factor (bFGF) to induce formation of fibrosis in
cHL. In a mouse model for fibrosis, the simultaneous application of TGF-β and bFGF causes
persistent fibrosis [24]. Both TGF-β and bFGF
are produced by HRS cells as well as the reactive background [25, 26]. They are produced
more prominently in nodular sclerosis than in
mixed cellularity cHL [27]. The third factor
that stimulates fibroblasts in cHL is the engagement of CD40. CD40, a member of the tumor
necrosis factor receptor (TNFR) superfamily,
can be upregulated on fibroblasts by IFN-γ. The
ligand of CD40 (CD40L) is present on activated
T cells, mast cells, and eosinophils present in
the cHL microenvironment.
4.1.7
osinophils, Plasma Cells,
E
Mast Cells, and B Cells
Presence of eosinophils in the reactive infiltrate
can be promoted by both IL-5, produced by Th2
cells, and by IL-9. In cHL patients with eosinophilia in the peripheral blood, HRS cells produce
IL-5 and IL-9 [28]. In addition, eosinophils are
attracted to cHL tissues by the production of the
chemokine CCL11, especially in nodular sclerosis cHL. CCL11 levels can be enhanced by the
production of tumor necrosis factor-α (TNF-α) by
the HRS cells, which in turn can induce CCL11
production in fibroblasts. This process is specific
for cHL since other lymphomas with tissue eosinophilia show no expression of CCL11 [29]. HRS
cells also produce CCL28 (MEC), and expression
of CCL28 correlates with the presence of eosinophils and plasma cells in cHL. CCL28 attracts
eosinophils by signaling through the chemokine
receptor CCR3 and attracts plasma cells through
CCR10 [30]. CCL5 is produced at high levels by
the reactive infiltrate in cHL and can attract eosinophils as well as mast cells. CCL5 and IL-9 may
both contribute to the attraction of mast cells in
cHL [31]. The stimulation and recruitment of
eosinophils in cHL can be illustrated in bone marrow biopsies that often show enhanced granulopoiesis with many eosinophils in the absence of
HRS cells. IL-6 produced by HRS cells in some
cases of cHL may explain the presence of variable
numbers of plasma cells [32]. B cells that express
CD20, CD21, IgM, and bcl-6 can be found in the
microenvironment of cHL [33]. It is possible that
these cells are remnants of the original lymph
node B cell areas. Plasma cells, mast cells, and
eosinophils are generally absent in NLPHL.
4.2
ross-Talk between HRS
C
Cells and Microenvironment
(Fig. 4.4)
4.2.1
Factors Supporting
Tumor Growth
It is likely that HL tumor cells originate from a
precursor B cell that has become addicted to
4
Microenvironment, Cross-Talk, and Immune Escape Mechanisms
75
Fig. 4.4 Schematic overview of the cross-talk between
Hodgkin/Reed-Sternberg (HRS) cells and the microenvironment. HRS cells attract specific subsets of cells by producing chemokines, are dependent on growth factors, and
use mechanisms of immunosuppression and immune
escape. Arrows indicate stimulating effects; the other lines
indicate inhibitory effects
activating and growth-supporting stimuli during
a deregulated immune response. Many additional events are needed to account for the highly
deregulated malignant phenotype of HRS and
LP cells. Although the tumor cells attain multiple alternative mechanisms to circumvent the
dependence on growth-stimulating signals from
the reactive infiltrate, they usually are not selfsufficient at the time of diagnosis. This is
reflected by the inability to generate cell lines
from primary HL cell suspensions.
IL-3 can function as a growth factor for B
cells and is produced by activated Th2 cells, mast
cells, and eosinophils. Its functions include protection against apoptosis and stimulation of proliferation. Most HRS cells express the IL-3
receptor, and exogenous IL-3 promotes growth of
cHL cell lines. Costimulation of HL cell lines
with IL-3 and IL-9 results in a further enhancement of cell growth [34]. There is no evidence
that HRS cells produce IL-3, so this signaling
pathway depends on IL-3 produced by the reactive infiltrate. In contrast, IL-7 is most likely an
autocrine as well as a paracrine growth factor for
HRS cells, since HRS cells express both the IL-7
receptor and produce IL-7 [35]. cHL cell lines
also produce IL-7, albeit at very low levels, and
anti-IL-7 treatment has some effect on cell
growth. Addition of IL-7 to HL cell line cultures
increased proliferation and protected against
apoptosis. Moreover, fibroblasts isolated from
cHL tissues are able to produce IL-7 [36]. Other
growth factors important for HRS cells are IL-9,
IL-13, IL-15, and, possibly, IL-6. IL-9 is
expressed by the tumor cells and not by the
infiltrating cells, and the IL-9 receptor is
­
expressed on cHL tumor cells and mast cells.
IL-9 supports tumor growth in cell lines and
functions as an autocrine factor in cHL tissue
[31]. IL-13 produced by HRS cells as well as the
L. Visser et al.
76
surrounding T cells drives proliferation and is
mostly autocrine [37]. Both IL-15 and the IL-15R
are expressed by HRS cells. IL-15 induces proliferation of HL cell lines and protects them against
apoptosis [38]. IL-6 is mainly produced by HRS
cells and occasionally by the infiltrating cells
[32]. In general, IL-6 is found at higher levels in
EBV-positive cases [39]. IL-6 might have an
autocrine effect although neutralizing antibodies
have no effect on the growth of cHL cell lines.
The signaling of cytokines upon binding to
their receptors leads to activation of the JAK-­
STAT pathway. This pathway is constitutively
activated in HL cell lines [40, 41], and several of
the STAT family members are expressed in HRS
cells of primary cHL cases [42, 43]. In addition
to the presence of cytokines, amplifications of
9p24.1, including the JAK2 gene locus found in
part of the cHL cases [44], can further enhance
constitutive activation of this pathway. Functional
studies in cHL cell lines have shown that STAT3
is involved in proliferation [45], while STAT1
and STAT6 play a role in protection against apoptosis [46]. Binding of IL-21 to the IL-21R
expressed on HL cell lines causes phosphorylation of STAT5 and induces proliferation [47].
HRS cells express several members of the
TNFR superfamily including CD30, which has
been used as a marker for cHL since the early
1980s. The CD30 ligand (CD30L) is expressed on
eosinophils [48] and mast cells [49] that are present
in the cHL infiltrate. Circulating eosinophils in
cHL patients also have increased expression levels
of CD30L [48]. Binding of CD30L to CD30 causes
enhanced secretion of IL-6, TNFα, lymphotoxin-α,
increased expression of ICAM-1 and B7, and, possibly, increased clonogenic growth and protection
against apoptosis in cHL cell lines [50]. Another
TNFR expressed on HRS cells is CD40. CD40 is
generally found on B cells, and B cells can be activated through CD40. In vitro rosetting of activated
CD4+ T cells around HRS cells is mediated
through the CD40L adhesion pathway [51].
Engagement of CD40 is important for the prevention of apoptosis. Similar to stimulation of CD30,
stimulation of HRS cell lines with CD40L causes
enhanced secretion of several cytokines and upregulation of costimulatory molecules [50].
Several receptor tyrosine kinases (RTKs) are
expressed by HRS cells and can have a role in
cell growth. Their ligands are expressed on cells
present in the microenvironment or by the HRS
cells themselves. Inhibition of PDGFRA,
expressed by the HL tumor cells, by imatinib
blocks proliferation. Its ligand, PDGFA, is also
produced by the HRS cells indicating autocrine
signaling [52]. DDR1 [53] and DDR2 [52] can
protect HRS cells from cell death by binding to
collagen, which is present in the immediate surrounding of HRS cells. Knockdown of DDR1
decreases survival of the L428 cHL cell line [53].
TRKA, the receptor for NGF, is expressed by
granulocytes [52], and TRK inhibition decreases
growth of cHL cell lines [54]. EPHB1 and its
ligand ephrin-B1 are both expressed by HRS
cells [52]. The HGF receptor c-Met is expressed
on HRS cells and inhibition causes G2/M cell
cycle arrest in HL cell lines. HGF is produced by
the tumor cells in a small group of patients and by
dendritic reticulum cells [55]. Insulin-like growth
factor receptor (IGF-1R) is expressed in 55% of
cHL patients, and inhibition of IGF-1R decreases
cell growth and induces G2/M cell cycle arrest in
HL cell lines [56]. Its ligand IGF-1 is expressed
by cells in the microenvironment [57]. PDGFRA,
DDR2, EPHB1, RON, TRKA, and TRKB are
found especially in EBV-negative HL [58], while
DDR1 is upregulated by LMP1 [53].
The Notch1 receptor is an upstream regulator
of NFκB [59]. It is highly expressed by HRS cells
and stimulation via Jagged1 induces proliferation
and survival of cHL cell lines [60].
4.2.2
Shaping the Environment
In addition to the production of several growth
factors, HRS cells also produce large amounts of
chemokines to attract specific beneficial or nonreacting cells. The lack of CD26 on the T cells
surrounding the HRS cells may result in an incapability to cleave chemokines and thereby modulates the chemotactic effects exerted by the HRS
cells. The attraction of specific populations of
cells is an important immune escape mechanism
exerted by the tumor cells.
4
Microenvironment, Cross-Talk, and Immune Escape Mechanisms
The most abundant and cHL-specific chemokine is CCL17 (TARC); it binds to CCR4 on Th2
cells, Treg cells, basophils, and monocytes.
CCL17 is highly expressed by HRS cells in ~95%
of cHL patients but not in NLPHL and most non-­
Hodgkin lymphomas [61, 62]. CCL17 can be
measured in serum and plasma and is a sensitive
and specific marker reflecting cHL tumor burden
[63–67]. High expression levels of CCL17 might
explain the influx of lymphocytes with a Th2and Treg-like phenotype, and CCL17-positive
cases are indeed associated with a higher percentage of CCR4-positive cells (Fig. 4.2; [62,
68]). In turn, Th2-type cytokines (IL-4, IL-13)
can induce the production of CCL17 by HRS
cells. CCR4-positive T cells are found especially
in the rosettes immediately surrounding the HRS
cells [15, 69]. CCL22 is another chemokine that
has a similar function as CCL17. High CCL22
protein expression levels were found in the cytoplasm of HRS cells in 90–100% of cHL patients
and also in tumor cells in the majority of NLPHL
and non-HL patients [70–73]. CCL22 production
can also be stimulated by Th2 cytokines, IL-4
and IL-13, and may reinforce the attraction of
Th2 and Treg cells, initiated by CCL17.
Stimulation of the IL-21 receptor on HRS by
IL-21 activates STAT3, which can induce CCL20
(MIP3α) production. CCL20 in turn attracts
memory T cells and Treg cells [74]. HRS cells
express both IL-21 and the IL-21 receptor, indicating presence of an autocrine signaling loop.
The expression of some chemokines is more pronounced in EBV-positive cHL (i.e., CXCL9 and
CXCL10), and as a result the composition of the
reactive background is somewhat different from
that in EBV-negative cHL, with a slightly higher
proportion of CD8+ T cells in EBV-positive cases
and more T cells with a Tr-1 phenotype (expressing LAG3, ITGA2, and ITGB2) [75]. T cell
recruitment is also enhanced by the upregulation
of adhesion molecules on endothelial cells,
induced by LTα [76] produced by HRS cells [77].
In addition to attracting specific cell subsets
by chemotaxis, HRS cells also shape their environment by inducing differentiation of specific T
cell subsets that are favorable for HRS cell survival and growth. The expression of IL-13 by the
77
HRS cells stimulates differentiation of naïve T
cells to Th2 cells [37]. The production of IL-7 by
HRS cells and fibroblasts can induce proliferation of Tregs [36]. Also, cHL cell lines with
antigen-­
presenting functions like KMH2 and
L428 have been shown to promote the differentiation of Treg like cells in vitro (expressing CD4,
CD25, FoxP3, CTLA4, and GITR and producing
large amounts of IL-10). Interestingly, these cell
lines can also induce the formation of CD4+
cytotoxic T cells (expressing granzyme B and
TIA-1) that can kill tumor cells directly, suggesting that CD4+ cytotoxic T cells have the potential
to attack tumor cells in vivo [78].
4.3
Immune Escape
Mechanisms (Fig. 4.4)
4.3.1
Antigen Presentation
The importance of antigen presentation in the
pathogenesis of cHL has been suggested by the
association of specific HLA subtypes with
increased cHL incidence. cHL is more common
in Caucasians as compared to Asians and about
4.5% of cHL cases occur in families [79, 80]. A
three- to sevenfold increased risk has been
observed in first-degree relatives and siblings. In
monozygotic twins, the co-twin has an approximate 100-fold increased risk of developing cHL
compared to dizygotic twins [81]. From the
1970s, a number of serological HLA types have
been associated with the occurrence of cHL. More
recently, a genetic screen of the entire HLA
region showed a strong association between the
HLA-A gene and EBV-positive cHL and the HLA
class II region with EBV-negative cHL [82, 83].
Four independent genome-wide association studies have confirmed that the HLA region is the
strongest genetic susceptibility locus in cHL [84–
87]. In EBV-positive cHL, it can be hypothesized
that this association is related to insufficient presentation of EBV antigenic peptides. These antigenic peptides most likely are derived from the
latency type II genes that are expressed in cHL,
i.e., LMP1, LMP2, and EBV-related nuclear antigen 1 (EBNA1). EBV partially escapes cytotoxic
L. Visser et al.
78
immune responses by downregulating immunodominant latent genes (EBNA2 and EBNA3). In
addition, the glycine-alanine repeat in EBNA1
largely prevents its presentation by HLA class I
by blocking its degradation into antigenic peptides through the proteasome [88]. However, subdominant immune responses to LMP2 and to a
lesser extent LMP1 are present in healthy EBV-­
infected individuals [89]. In fact, adoptive immunotherapy in relapsed EBV-positive cHL has
been used in some small studies with success. In
these studies, peripheral blood from cHL patients
was used to generate EBV-specific cytotoxic T
cells in vitro, and these were reinfused. Some
durable complete responses were observed, with
better responses if the cytotoxic T cells were specifically targeted to LMP2 [90–92] (Fig. 4.4).
Interestingly, the genetic association of the HLA-­
A gene with EBV-positive cHL is attributed to the
presence of the HLA-A∗01 type and absence of
the HLA-A∗02 type [93]. HLA-A∗01 is known
to have a low affinity for LMP2- and LMP1-­
derived antigenic peptides, while HLA-A∗02 can
present these peptides very well. This suggests
that EBV-positive cHL is more likely to occur
after primary EBV infection if an individual’s set
of HLA class I molecules cannot properly present LMP2 and LMP1 to the immune system [94].
4.3.2
HLA Class I Expression
Defects in the antigen-presenting pathways are
very common in solid malignancies, as well as in
many B cell lymphomas, and are an obvious
mechanism to escape from antitumor immune
responses. In EBV-negative cHL, less than 20%
of cases retain expression of cell surface HLA
class I on the HRS cells at the time of diagnosis.
Paradoxically, HLA class I expression by HRS
cells is retained in ~75% of EBV-positive cHL
patients [95–97]. One common mechanism of
HLA class I loss is presence of somatic mutations
in the β2-microglobulin gene. This leads to loss
of β2-microglobulin protein, which is necessary
for HLA class I assembly and transport to the cell
surface. Other mechanisms also appear to be
involved as immunohistochemistry has shown
cytoplasmic β2-microglobulin expression in part
of the cases that lost HLA class I heavy chain
expression [98]. These different mechanisms
may indicate that downregulation of HLA class I
is based on clonal selection by continuous cytotoxic immune responses. This may be related to
the presence of antigenic peptides that are related
to malignant transformation or disease progression. However, downregulation of HLA class I
generally induces activation of natural killer
(NK) cells. These cells contain HLA class
I-specific inhibitory receptors and are sparse in
the reactive infiltrate of cHL. The inhibitory
receptors can also be engaged by the nonclassical
HLA class I-like molecule known as HLA-G. In
about two thirds of the HLA class I-negative cHL
cases, the HRS cells indeed express HLA-G [98].
Besides NK cell inhibition, HLA-G might also
induce Treg cells and inhibit cytotoxic T cell
responses. Another immune escape mechanism
consists of the proteolytic cleavage of MHC
(HLA) class I-related chain-A (MIC-A) by ERp5
and ADAM10, which are both expressed by HRS
cells. MIC-A is a membranous ligand for the activating NKG2D receptor present on cytotoxic T
cells. In addition, the NKG2D receptor expression by these cytotoxic T cells is reduced in the
presence of TGF-β [99].
4.3.3
HLA Class II Expression
HLA class II cell surface expression on HRS
cells is lost in approximately 40% of all cHL
patients [95]. In addition, translocations involving CIITA have been found in 15% of cHL
patients and may result in partial downregulation
of HLA class II expression [101]. The absence of
HLA class II is weakly related to extranodal disease, EBV-negative status, and absence of HLA
class I cell surface expression. Lack of HLA class
II expression has been associated with adverse
failure-free survival and relative survival and is
independent of other prognostic factors [95]. It
can be hypothesized that antigen presentation in
the context of HLA class II is involved in recruitment and activation of CD4+ T cells early in cHL
pathogenesis. Under the influence of immunomodulating mechanisms, these T cells are important in providing trophic factors for HRS cells
4
Microenvironment, Cross-Talk, and Immune Escape Mechanisms
and also have a role in inhibiting Th1 responses.
In the initial stages of cHL pathogenesis, HRS
cells are probably highly dependent on the reactive infiltrate and expression of HLA class II, but
as the lymphoma develops, this dependency may
weaken because of alternative trophic and immunosuppressive strategies. Thus, downregulation
of HLA class II without loss of viability of HRS
cells might occur when the HRS cells have grown
less dependent on the reactive infiltrate. This is
supported by the finding that downregulation of
HLA class II is associated with extranodal disease [95].
4.3.4
Immune Checkpoints
Immune checkpoint molecules have gained much
attention due to their use as treatment targets.
Both CTLA-4 and PD-1 blockade have shown
remarkable results in cHL patients in clinical
trials.
CTLA-4 is expressed exclusively on T cells
upon activation. It gives an inhibitory signal early
after T cell activation, by binding to CD80/CD86
with a higher avidity than the costimulatory molecule CD28. This limits T cell activation and proliferation [101, 102]. Interestingly, CTLA-4 is
present on the characteristic CD26− T cells in
HL [15]. Moreover, HL cell lines are able to
induce differentiation of naïve T cells into
CTLA4+ Tregs [78]. So far, two clinical trials
have exploited the use of a monoclonal antibody
targeting CTLA-4 in relapsed and refractory HL
after allogeneic hematopoietic cell transplantation. Objective response rates were observed in 2
out of 14 and 1 out of 7 cases [103, 104].
The interaction partner of PD-L1, PD-1, is
present on activated T cells, B cells, macrophages
(which also express PD-L1), and NK cells within
the microenvironment [105, 106]. PD-L1 is highly
expressed on HRS cells [105, 107], due to a selective amplification of the PD-L1 region on 9p24.1
[44], activation of AP-1 and LMP-1 [107], or
chromosomal alterations involving the CIITA
locus [100]. Anti-PD-1 therapy in relapsed and
refractory HL patients, using the monoclonal antibodies nivolumab or pembrolizumab, showed
objective response rates in 65–87% of the cHL
79
patients [108–111]. The mechanism of action of
PD-1 blockade in HL remains unknown, but multiple mechanisms have been studied. In contrast to
solid tumors where CD8+ cytotoxic T cells seem
to be the main effector cells [112], CD4+ T cells
might have an important role in mediating the
antitumor immune response in HL. CD8+ T cells
recognize the tumor cells through (neo)antigens
presented in the context of HLA class I, which
can ultimately lead to eradication of the tumor
cells. However, HLA class I is often absent on
HRS cells, making a central role for CD8+ T cells
in immune checkpoint efficacy unlikely [96]. In
HL, the inflammatory infiltrate is mainly dominated by CD4+ T cells, which are more often in
direct contact with the HRS cells when compared
to CD8+ T cells. In addition, CD4+PD-1+ T cells
are more frequently bound to PD-L1+ HRS cells
[113]. The majority of complete responders to
nivolumab lack membranous HLA class I expression, while being positive for membranous HLA
class II expression. Also, presence of HLA class II
is predictive for a prolonged progression-free survival in patients treated with nivolumab
>12 months after myeloablative autologous stem
cell transplantation, in contrast to presence of
HLA class I [114]. Although many studies on
PD-1 blockade focused on T cells, expression of
PD-1 on T cells in direct contact with HRS cells is
rare, and their numbers are significantly lower in
cases with PD-L1 gain [115]. Interestingly,
exhausted PD-1+ CD3+CD56hiCD16− NK cells
are enriched in HL, and their activation can be
inhibited by PD-L1+CD163+CD14+ tumor-­
associated macrophages, which are also increased
in HL. This inhibition was effectively reversed by
PD-1 blockade [106]. This indicates the
­importance of other cell types in responses to
PD-1 blockade that are currently less well
characterized.
An interesting molecule with regard to the role
of CD4+ T cells in responses to anti-PD-1 therapy
is LAG-3. LAG-3 is an immune checkpoint molecule expressed on activated T cells, NK cells, B
cells, and plasmacytoid dendritic cells [116]. The
main interaction partner for LAG-3 is HLA class
II, to which LAG-3 binds with higher affinity than
CD4 [117]. LAG-3 is upregulated on Tregs, and
LAG-3-positive lymphocytes are enriched in the
80
proximity of HRS cells [118, 119]. Increased
LAG-3 expression was observed especially in
EBV-positive cHL cases [118, 120]. Interestingly,
the percentage of LAG-3-positive cells was
enriched in CD4+ T cells that express a high intracellular CTLA-4 and GITR, but not FOXP3+
[118]. The GITRhi CD4+ T cells are frequently in
direct contact with HRS cells [15]. Moreover,
CD4+LAG-3+ cells are significantly expanded in
patients with active disease, which is in concordance with the ability of HL cell line supernatant
to increase expansion of CD4+LAG-3+ regulatory T cells within PBMCs. In addition, higher
FOXP3 and LAG-3 expression on tumor-infiltrating lymphocytes is associated with decreased
LMP1- and LMP2-­specific CD8+ T cell function
[118]. Altogether this implicates an important role
for LAG-3 in inhibiting (EBV) mediated cellular
immunity in HL and points to LAG-3 as an interesting treatment target in combination with PD-1
blockade.
Currently, more and more immune checkpoint
molecules are emerging as potential targets for
therapy. Some of these are less well studied in the
context of HL. For example, HRS cells express
HVEM and CD200/CD200R [121] in addition to
the earlier mentioned checkpoints, whereas TIM-­
3+ T cells are present within the inflammatory
infiltrate [120], indicating the complexity of
immunosuppressive mechanisms within the HL
microenvironment.
L. Visser et al.
nosuppressive cytokines TGF-β and IL-10.
TGF-β is produced by HRS cells in nodular sclerosis cHL [25, 26], whereas IL-10 is more frequently found in EBV-positive (mixed cellularity)
cHL [122, 123]. In normal cells, TGF-β is produced in an inactive form, which can be activated
by acidification. TGF-β produced by cHL cell
lines is active at a physiological pH and has a
high molecular weight [124]. The same high
molecular weight form of TGF-β can also be
found in the urine of cHL patients [125] indicating that in patients HRS cells are able to produce
the active TGF-β form.
Tregs present in the microenvironment of cHL
are highly immunosuppressive and contain Tr1
(IL-10-producing Tregs) as well as CD4+CD25+
Tregs. IL-10, cell-cell contact, and CTLA4 play a
main role in executing their immunosuppressive
function [126]. In addition, HRS cells express
galectin-1, an animal lectin, which can cause
apoptosis in activated T cells, induce differentiation into Treg cells, and contribute to the elimination of an effective antitumor response in cHL
[127]. Galectin-1 expression blocks CD8+ T cell
responses against LMP1 and LMP2 in EBV-­
positive cHL [128]. HRS cells express FAS and
the FAS ligand. However, there are some mechanisms protecting the HRS cells from apoptosis
induction, such as FAS mutations in a small proportion of cases and c-FLIP overexpression in all
cases [129]. Presumably, activated Th1 and
CD8+ T cells expressing FAS are driven into
apoptosis by the FAS ligand expression on the
4.3.5 Immunosuppression
HRS cells. Also, indoleamine 2,3-dioxygenase
(IDO), a known immunosuppressor, is expressed
As normal B cells are professional antigen-­ by histiocytes, dendritic cells, and endothelial
presenting cells, HRS cells are expected to pres- cells in the microenvironment of cHL. IDO is
ent antigens to the immune system, at least early found more often in EBV-positive HL, upreguin disease pathogenesis. Indeed, most compo- lates the number of Tregs [130], and potentially
nents of the HLA class I and HLA class II blocks the CD8+ T cell response [131]. In EBV-­
antigen-­presenting pathways have been detected positive cHL, the Th1-inducing cytokine IL-12 is
in the HRS cells at the time of diagnosis. expressed in T cells surrounding the HRS cells,
However, Th1 cells are not actively attracted by and its presence suggests that these T cells have
the HRS cells and CD8+ T cells are relatively the potential to induce antitumor activity [132].
scarce. Moreover, HRS cells have gained the However, an EBV-induced IL-12-related cytocapacity to prevent CD8+ T cells from attacking kine called EBI3 can block this Th1 response and
by producing high amounts of the strongly immu- is produced by HRS cells [133].
4
Microenvironment, Cross-Talk, and Immune Escape Mechanisms
4.4
Prognostic Impact
of the Microenvironment
Several research groups studied the cHL reactive infiltrate in relation to prognosis. Gene
expression profiling of whole tissue and subsequent validation by immunohistochemistry
showed that high numbers of CD68-positive
cells are related to adverse outcome [134]. In a
meta-analysis of almost 3000 patients, a high
density of CD68+ tumor-associated macrophages predicted overall survival, shorter progression-free survival, and poor disease survival
in adult cHL [135]. Similar findings were
obtained for CD163+ macrophages in this study.
Analysis of 100 pediatric cHL revealed that
especially M2 macrophages, characterized by
co-expression of CD163 and c-Maf, are associated with poor survival, while M1 macrophages
are associated with better survival [136]. In
some studies tumor-associated macrophages
were associated with EBV-positive tumors [135,
137] and presence of other cell types in the
microenvironment, such as cytotoxic T cells
[136] and mast cells [138]. Patients with a
higher degree of mast cell infiltration or with
tissue eosinophilia have an adverse failure-free
survival, probably because the CD30L expression by these cell types is advantageous to the
HRS cells [48, 49, 138].
Large numbers of Th2 cells in the microenvironment, as determined by c-Maf expression, correlate with improved disease-free survival [14].
Also, increased numbers of infiltrating Treg cells
seem to correlate with improved survival as this
effect was observed in two out of three studies
[14, 139, 140]. Accordingly, a high percentage of
activated CD8+ granzyme B+ T cells is a strong
indicator of unfavorable clinical outcome [141].
More recently, a high proportion of Treg cells and
the associated anergic phenotype of the microenvironment has been associated with a shorter time
to progression [137]. A high CD4/CD8 T cell
ratio was associated with treatment failure [142].
A high ratio of FoxP3 to cytotoxicity markers
granzyme B [140] or Tia-1 [139] gives the best
predictive value for a good prognosis and has also
been correlated to the presence of macrophages
81
[136, 138], which might—in part—explain these
effects. In other malignancies the presence of
Tregs and the absence of CD8+ T cells have been
associated with adverse prognosis. One explanation of this opposite effect might be that HRS
cells are expected to behave more aggressively as
they develop a stronger independency from the
reactive infiltrate. In this situation a hostile microenvironment is allowed, because the HRS cells
have acquired alternative immune evasive strategies. This theory fits with the adverse prognostic
impact of absence of HLA class II expression.
4.5
Conclusion
The microenvironment is a fundamental component of the tumor mass and an essential
pathogenetic factor in cHL and NLPHL. It supplies the tumor cells with growth factors and
inhibits antitumor immune responses. In fact, it
could be stated that the infltrate does not consist
of ‘innocent bystanders’ but contains ‘guilty
opportunists’ [31]. As the tumor cells and the
reactive infiltrate grow up together, there is an
extensive cross-talk between these two components. The tumor cells actively attract and shape
their environment for their own benefit and
make use of a number of mechanisms to fend
off antitumor immune responses.
References
1. Poppema S, Delsol G, Pileri SA et al (2008) Nodular
lymphocyte predominant Hodgkin lymphoma. In:
Swerdlow SH, Campo E, Harris NL et al (eds) WHO
classification of tumours of haematopoietic and lymphoid tissues. IARC, Lyon
2. Stein H, Delsol G, Pileri SA et al (2016) Classical
Hodgkin lymphoma, introduction. In: Swerdlow SH,
Campo E, Harris NL et al (eds) WHO classification
of tumours of haematopoietic and lymphoid tissues.
IARC, Lyon
3. Fan Z, Natkunam Y, Bair E, Tibshirani R, Warnke
RA (2003) Characterization of variant patterns of
nodular lymphocyte predominant Hodgkin lymphoma with immunohistologic and clinical correlation. Am J Surg Pathol 27:1346–1356
4. Atayar C, van den Berg A, Blokzijl T et al (2007)
Hodgkin lymphoma associated T-cells exhibit a
L. Visser et al.
82
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
transcription factor profile consistent with distinct lymphoid compartments. J Clin Pathol
60:1092–1097
Carbone A, Gloghini A, Cabras A et al (2009)
Differentiating germinal center-derived lymphomas through their cellular microenvironment. Am J
Hematol 84:435–438
Sattarzadeh A, Visser L, Rutgers B, Diepstra A, van
den Berg A (2016) Characterization of the microenvironment of nodular lymphocyte-predominant
Hodgkin lymphoma. Int J Mol Sci 17:e2127
Huppmann AR, Nicolae A, Slack GW et al (2014)
EBV may be expressed in the LP cells of nodular lymphocyte-predominant Hodgkin lymphoma
(NLPHL) in both children and adults. Am J Surg
Pathol 38:316–324
Oudejans JJ, Jiwa NM, Kummer JA et al (1996)
Analysis of major histocompatibility complex class
I expression on Reed-Sternberg cells in relation to
the cytotoxic T-cell response in Epstein-Barr virus-­
positive and -negative Hodgkin’s disease. Blood
87:3844–3851
Goedert JJ, Cote TR, Virgo P et al (1998) Spectrum
of AIDS-associated malignant disorders. Lancet
351:1833–1839
Biggar RJ, Jaffe ES, Goedert JJ et al (2006) Hodgkin
lymphoma and immunodeficiency in persons with
HIV/AIDS. Blood 108:3786–3791
Poppema S (1996) Immunology of Hodgkin’s disease. Ballieres Clin Haematol 9:447–457
Wolf M, Albrecht S, Marki C (2008) Proteolytic processing of chemokines: implications in physiological and pathological conditions. Int J Biochem Cell
Biol 40:1185–1198
Von Bonin A, Huhn J, Fleischer B (1998) Dipeptidyl-­
peptidase IV/CD26 on T cells: analysis of an alternative T-cell activation pathway. Immunol Rev
161:43–53
Schreck S, Friebel D, Buettner M et al (2009)
Prognostic impact of tumour-infiltrating Th2 and
regulatory cells in classical Hodgkin lymphoma.
Hematol Oncol 27:31–39
Ma Y, Visser L, Blokzijl T et al (2008) The
CD4+CD26− T-cell population in classical
Hodgkin’s lymphoma displays a distinctive regulatory T-cell profile. Lab Investig 88:482–490
Greaves P, Clear A, Owen A et al (2013) Defining
characteristics of classical Hodgkin lymphoma microenvironment T helper cells. Blood
122:2856–2863
Wu R, Sattarzadeh A, Rutgers B, Diepstra A, van
den Berg A, Visser L (2016) The microenvironment
of classical Hodgkin lymphoma: heterogeneity by
Epstein-Barr virus presence and location within the
tumor. Blood Cancer J 6:e417
Cader FZ, Schackmann RCJ, Hu X et al (2018) Mass
cytometry of Hodgkin lymphoma reveals a CD4(+)
regulatory T-cell-rich and exhausted T-effector
microenvironment. Blood 132:825–836
19. Nam-Cha SH, Roncador G, Sanchez-Verde L
et al (2008) PD-1, a follicular T-cell marker useful for recognizing nodular lymphocyte-predominant Hodgkin lymphoma. Am J Surg Pathol
32:1252–1257
20. Sattarzadeh A, Diepstra A, Rutgers B, van den
Berg A, Visser L (2015) CD57+ T-cells are a subpopulation of T-follicular helper cells in nodular
lymphocyte-predominant Hodgkin lymphoma. Exp
Hematol Oncol 4:27
21. Atayar C, Poppema S, Visser L et al (2006) Cytokine
gene expression profile distinguishes CD4+/CD57+
T-cells of nodular lymphocyte predominance type
of Hodgkin lymphoma from their tonsillar counterparts. J Pathol 208:423–430
22. Rahemtullah A, Harris NL, Dorn ME et al (2008)
Beyond the lymphocyte predominant cell:
CD4+CD8+ T-cells in nodular lymphocyte predominant Hodgkin lymphoma. Leuk Lymphoma
49:1870–1878
23. Ohshima K, Akaiwa M, Umeshita R et al (2001)
Interleukin-13 and interleukin-13 receptor in
Hodgkin’s disease: possible autocrine mechanism and involvement in fibrosis. Histopathology
38:368–375
24. Shinozaki M, Kawara S, Hayashi N et al (1997)
Induction of subcutaneous tissue fibrosis in newborn
mice by transforming growth factor-b – simultaneous application with basic growth factor causes
persistent fibrosis. Biochem Biophys Res Commun
237:292–296
25. Kadin M, Butmarc J, Elovic A et al (1993)
Eosinophils are the major source of transforming
growth factor-beta 1 in nodular sclerosing Hodgkin’s
disease. Am J Pathol 142:11–16
26. Newcom SR, Gu L (1995) Transforming growth factor beta 1 messenger RNA in Reed-Sternberg cells in
nodular sclerosing Hodgkin’s disease. J Clin Pathol
48:160–163
27. Ohshima K, Sugihara M, Suzumiya J et al (1999)
Basic fibroblast growth factor and fibrosis in
Hodgkin’s disease. Pathol Res Pract 195:149–155
28. Samoszuk M, Nansen L (1990) Detection of interleukin-­5 messenger RNA in Reed-Sternberg cells
of Hodgkin’s disease with eosinophilia. Blood
75:13–16
29. Jundt F, Anagnostopoulos I, Bommert K et al (1999)
Hodgkin/Reed-Sternberg cells induce fibroblasts to
secrete eotaxin, a potent chemoattractant for T cells
and eosinophils. Blood 94:2065–2071
30. Hanamoto H, Nakayama T, Miyazato H (2004)
Expression of CCL28 by Reed-Sternberg cells
defines a major subtype of classical Hodgkin’s disease with frequent infiltration of eosinophils and/or
plasma cells. Am J Pathol 164:997–1006
31. Glimelius I, Edstrom A, Amini RM et al (2006)
IL-9 expression contributes to the cellular composition in Hodgkin lymphoma. Eur J Haematol
76:278–283
4
Microenvironment, Cross-Talk, and Immune Escape Mechanisms
32. Jucker M, Abts H, Li W et al (1991) Expression of
interleukin-6 and interleukin-6 receptor in Hodgkin’s
disease. Blood 77:2413–2418
33. Tudor CS, Distel LV, Eckhardt J et al (2013) B cells
in classical Hodgkin lymphoma are important actors
rather than bystanders in the local immune reaction.
Hum Pathol 44:2475–2486
34. Aldinucci D, Poletto D, Nanni P et al (2002)
Expression of functional interleukin-3 receptors on
Hodgkin and Reed-Sternberg cells. Am J Pathol
160:585–596
35. Foss HD, Hummel M, Gottstein S et al (1995)
Frequent expression of IL-7 gene transcripts in
tumor cells of classical Hodgkin’s disease. Am J
Pathol 146:33–39
36. Cattaruzza L, Gloghini A, Olivo K et al (2009)
Functional coexpression of interleukin (IL)-7 and
its receptor (IL-7R) on Hodgkin and Reed-Sternberg
cells: involvement of IL-7 in tumor cell growth and
microenvironmental interactions of Hodgkin’s lymphoma. Int J Cancer 125:1092–1101
37. Kapp U, Yeh WC, Patterson B et al (1999)
Interleukin 13 is secreted by and stimulates the
growth of Hodgkin and Reed-Sternberg cells. J Exp
Med 189:1939–1946
38. Ullrich K, Blumenthal-Barby F, Lamprecht B et al
(2015) The IL-15 cytokine system provides growth
and survival signals in Hodgkin lymphoma and
enhances the inflammatory phenotype of HRS cells.
Leukemia 29:1213–1218
39. Herbst H, Samol J, Foss HD et al (1997) Modulation
of interleukin-6 expression in Hodgkin and Reed-­
Sternberg cells by Epstein–Barr virus. J Pathol
182:299–306
40. Cochet O, Frelin C, Peyron J-F, Imbert V (2006)
Constitutive activation of STAT proteins in the
HDLM-2 and L540 Hodgkin lymphoma-derived cell
lines supports cell survival. Cell Signal 18:449–455
41. Kube D, Holtick U, Vockerodt M et al (2001) STAT3
is constitutively activated in Hodgkin cell lines.
Blood 98:762–770
42. Hinz M, Lemke P, Anagnostopoulos I, Hacker C
et al (2002) Nuclear factor kappaB-dependent gene
expression profiling of Hodgkin's disease tumor
cells, pathogenetic significance, and link to constitutive signal transducer and activator of transcription
5a activity. J Exp Med 196:605–617
43. Skinnider BF, Elia AJ, Gascoyne RD et al (2002)
Signal transducer and activator of transcription 6 is
frequently activated in Hodgkin and Reed-Sternberg
cells of Hodgkin lymphoma. Blood 99:618–626
44. Green MR, Rodig S, Juszczynski P et al (2012)
Constitutive AP-1 activity and EBV infection induce
PD-L1 in Hodgkin lymphomas and posttransplant
lymphoproliferative disorders; implications for targeted therapy. Clin Cancer Res 18:1611–1618
45. Baus D, Pfitzner E (2006) Specific function of
STAT3, SOCS1, and SOCS3 in the regulation of
proliferation and survival of classical Hodgkin lymphoma cells. Int J Cancer 118:1404–1413
83
46. Baus D, Nonnenmacher F, Jankowski S et al (2009)
STAT6 and STAT1 are essential antagonistic regulators of cell survival in classical Hodgkin lymphoma
cell line. Leukemia 23:1885–1893
47. Scheeren FA, Diehl SA, Smit LA et al (2008) IL-21
is expressed in Hodgkin lymphoma and activates
STAT5: evidence that activated STAT5 is required for
Hodgkin lymphomagenesis. Blood 111:4709–4715
48. Pinto A, Aldinucci D, Gloghini A et al (1996)
Human eosinophils express functional CD30 ligand
and stimulate proliferation of a Hodgkin’s disease
cell line. Blood 88:3299–3305
49. Molin D, Edstrom A, Glimelius I et al (2002) Mast
cell infiltration correlates with poor prognosis in
Hodgkin’s lymphoma. Br J Haematol 119:122–124
50. Grüss HJ, Ulrich D, Braddy S et al (1995)
Recombinant CD30 ligand and CD40 ligand share
common biological activities on Hodgkin and Reed-­
Sternberg cells. Eur J Immunol 25:2083–2089
51. Carbone A, Gloghini A, Gattei V et al (1995)
Expression of functional CD40 antigen on Reed-­
Sternberg cells and Hodgkin’s disease cell lines.
Blood 85:780–789
52. Renné C, Willenbrock K, Küppers R et al (2005)
Autocrine- and paracrine-activated receptor tyrosine kinases in classic Hodgkin lymphoma. Blood
105:4051–4059
53. Cader FZ, Vockerodt M, Bose S et al (2013) The
EBV oncogene LMP1 protects lymphoma cells from
cell death through the collagen-mediated activation
of DDR1. Blood 122:4237–4245
54. Renné C, Minner S, Küppers R et al (2008)
Autocrine NGFβ/TRKA signaling is an important
survival factor for Hodgkin lymphoma derived cell
lines. Leukemia Res 32:163–167
55. Xu C, Plattel W, van den Berg A et al (2012)
Expression of the c-met oncogene by tumor cells
predicts a favorable outcome in classical Hodgkin’s
lymphoma. Heamatologica 97:572–578
56. Liang Z, Diepstra A, Xu C et al (2014) Insulin-like
growth factor 1 receptor is a prognostic factor in
classical Hodgkin lymphoma. PLOSone 9:e87474
57. Eppler E, Janas E, Link K et al (2015) Insulin-like
growth factor I is expressed in classical and ­nodular
lymphocyte-predominant Hodgkin’s lymphoma
tumour and microenvironmental cells. Cell Tissue
Res 359:841–851
58. Renné C, Hinsch N, Willenbrock K et al (2007) The
aberrant coexpression of several receptor tyrosine
kinases is largely restricted to EBV-negative cases
of classical Hodgkin’s lymphoma. Int J Cancer
120:2504–2509
59. Schwarzer R, Dörken B, Jundt F (2012) Notch is an
essential upstream regulator of NF-κB and is relevant for the survival of Hodgkin and Reed-Sternberg
cells. Leukemia 26:806–813
60. Jundt F, Anagnostopoulos I, Förster R et al (2002)
Activated Notch1 signaling promotes tumor cell
proliferation and survival in Hodgkin and anaplastic
large cell lymphoma. Blood 99:3398–3403
L. Visser et al.
84
61. Peh SC, Kim LH, Poppema S (2001) TARC, a
CC chemokine, is frequently expressed in classic
Hodgkin lymphoma but not in NLP Hodgkin lymphoma, T-cell-rich B-cell lymphoma, and most cases
of anaplastic large cell lymphoma. Am J Surg Pathol
25:925–929
62. Van den Berg A, Visser L, Poppema S (1999) High
expression of the CC chemokine TARC in Reed-­
Sternberg cells. A possible explanation for the characteristic T-cell infiltrate in Hodgkin’s lymphoma.
Am J Pathol 154:1685–1691
63. Niens M, Visser L, Nolte IM et al (2008) Serum
chemokine levels in Hodgkin lymphoma patients:
highly increased levels of CCL17 and CCL22. Br J
Haematol 140:527–536
64. Weihrauch MR, Manzke O, Beyer M et al (2005)
Elevated levels of CC thymus and activation-related
chemokine (TARC) in primary Hodgkin’s disease: potential for a prognostic factor. Cancer Res
65:5516–5519
65. Plattel WJ, van den Berg A, Visser L et al (2012)
Plasma thymus and activation-regulated chemokine
as an early response marker in classical Hodgkin’s
lymphoma. Haematologica 97:410–415
66. Sauer M, Plütschow A, Jachimowicz RD et al (2013)
Baseline serum TARC levels predict therapy outcome in patients with Hodgkin lymphoma. Am J
Hematol 88:113–115
67. Plattel WJ, Alsada ZN, van Imhoff GW, Diepstra
A, van den Berg A, Visser L (2016) Biomarkers
for evaluation of treatment response in classical
Hodgkin lymphoma: comparison of sGalectin-1,
sCD163 and sCD30 with TARC. Br J Haematol
175(5):868–875
68. Ohshima K, Tutiya T, Yamaguchi T et al (2002)
Infiltration of Th1 and Th2 lymphocytes around
Hodgkin and Reed-Sternberg (H&RS) cells in
Hodgkin disease: relation with expression of CXC
and CC chemokines on H&RS cells. Int J Cancer
98:567–572
69. Ishida T, Ishii T, Inagaki A et al (2006) Specific
recruitment of CC chemokine receptor 4-positive
regulatory T cells in Hodgkin lymphoma fosters
immune privilege. Cancer Res 66:5716–5722
70. Andrew DP, Chang MS, McNinch J et al (1998)
STPC-1 (MDC) CC chemokine acts specifically on
chronically activated Th2 lymphocytes and is produced by monocytes on stimulation with Th2 cytokines IL-4 and IL-13. J Immunol 16:5027–5038
71. Hedvat CV, Jaffe ES, Qin J et al (2001) Macrophage-­
derived chemokine expression in classical Hodgkin’s
lymphoma: application of tissue microarrays. Mod
Pathol 14:1270–1276
72. Imai T, Chantry D, Raport CJ et al (1998)
Macrophage-derived chemokine is a functional
ligand for the CC chemokine receptor 4. J Biol
Chem 273:1764–1768
73. Maggio E, van den Berg A, Visser L et al (2002)
Common and differential chemokine expression patterns in RS cells of NLP, EBV positive
74.
75.
76.
77.
78.
79.
80.
81.
82.
83.
84.
85.
86.
and negative classical Hodgkin lymphomas. Int J
Cancer 99:665–672
Lamprecht B, Kreher S, Anagnostopoulos I et al
(2008) Aberrant expression of the Th2 cytokine
IL-21 in Hodgkin lymphoma cells regulates STAT3
signaling and attracts Treg cells via regulation of
MIP-3alpha. Blood 112:3339–3347
Morales O, Mrizak D, François V et al (2014)
Epstein-Barr virus infection induces an increase
of T regulatory type 1 cells in Hodgkin lymphoma
patients. Br J Haematol 166:875–890
Fhu CW, Graham AM, Yap CT et al (2014)
Reed-­
Sternberg cell-derived lymphotoxin-alpha
activates endothelial cells to enhance T-cell recruitment in classical Hodgkin lymphoma. Blood
124:2973–2982
Foss H-D, Herbst H, Oelmann E et al (1993)
Lymphotoxin, tumour necrosis factor and interleukin-­6 gene transcripts are present in Hodgkin and
Reed-Sternberg cells of most Hodgkin’s disease
cases. Br J Haematol 84:627–635
Tanijiri T, Shimizu T, Uehira K et al (2007) Hodgkin’s
Reed-Sternberg cell line (KM-H2) promotes a bidirectional differentiation of CD4+CD25+Foxp3+
T cells and CD4+ cytotoxic T lymphocytes from
CD4+ naive T cells. J Leukoc Biol 82:576–584
Ferraris AM, Racchi O, Rapezzi D et al (1997)
Familial Hodgkin’s disease: a disease of young
adulthood? Ann Hematol 74:131–134
Glaser SL, Hsu JL (2002) Hodgkin’s disease in
Asians: incidence patterns and risk factors in
population-­based data. Leuk Res 26:261–269
Mack TM, Cozen W, Shibata DK et al (1995)
Concordance for Hodgkin’s disease in identical
twins suggesting genetic susceptibility to the young-­
adult form of the disease. N Engl J Med 332:413–418
Diepstra A, Niens M, Vellenga E et al (2005)
Association with HLA class I in Epstein–Barr-­
virus-­positive and with HLA class III in Epstein-­
Barr-­
virus-negative Hodgkin’s lymphoma. Lancet
365:2216–2224
Niens M, van den Berg A, Diepstra A et al (2006)
The human leukocyte antigen class I region is associated with EBV-positive Hodgkin’s lymphoma:
HLA-A and HLA complex group 9 are putative
candidate genes. Cancer Epidemiol Biomark Prev
15:2280–2284
Enciso-Mora V, Broderick PMY et al (2010) A
genome-wide association study of Hodgkin’s lymphoma identifies new susceptibility loci at 2p16.1
(REL), 8q24.21 and 10p14 (GATA3). Nat Genet
42:1126–1130
Urayama KY, Jarrett RF, Hjalgrim H et al (2012)
Genome-wide association study of classical
Hodgkin lymphoma and Epstein-Barr virus status-­
defined subgroups. J Natl Cancer Inst 104:240–253
Cozen W, Li D, Best T et al (2012) A genome-­
wide meta-analysis of nodular sclerosing Hodgkin
lymphoma identifies risk loci at 6p21.32. Blood
119:469–475
4
Microenvironment, Cross-Talk, and Immune Escape Mechanisms
87. Sud A, Thomsen H, Orlando G et al (2018) Genome-­
wide association study implicates immune dysfunction in the development of Hodgkin lymphoma.
Blood 132:1212–1218
88. Levitskaya J, Coram M, Levitsky V et al (1995)
Inhibition of antigen processing by the internal
repeat region of the Epstein–Barr virus nuclear antigen-­1. Nature 375:685–688
89. Meij P, Leen A, Rickinson AB et al (2002)
Identification and prevalence of CD8(+) T-cell
responses directed against Epstein–Barr virus-­
encoded latent membrane protein 1 and latent membrane protein 2. Int J Cancer 99:93–99
90. Bollard CM, Aguilar L, Straathof KC et al (2004)
Cytotoxic T lymphocyte therapy for EpsteinBarr virus+ Hodgkin’s disease. J Exp Med
200:1623–1633
91. Lucas KG, Salzman D, Garcia A et al (2004)
Adoptive immunotherapy with allogeneic Epstein–
Barr virus (EBV)-specific cytotoxic T-lymphocytes
for recurrent EBV-positive Hodgkin disease. Cancer
100:1892–1901
92. Bollard CM, Gottschalk S, Torrano V et al (2014)
Sustained complete responses in patients with lymphoma receiving autologous cytotoxic T lymphocytes targeting Epstein-Barr virus latent membrane
proteins. J Clin Oncol 32:798–808
93. Niens M, Jarrett RF, Hepkema B et al (2007) HLA-­
A∗02 is associated with a reduced risk and HLA-­
A∗01 with an increased risk of developing EBV+
Hodgkin lymphoma. Blood 110:3310–3315
94. Jones K, Wockner L, Brennan RM et al (2016)
The impact of HLA class I and EBV latency-II
antigen-specific CD8+ T cells on the pathogenesis
of EBV+ Hodgkin lymphoma. Clin Exp Immunol
183:206–220
95. Diepstra A, van Imhoff GW, Karim-Kos HE et al
(2007) HLA class II expression by Hodgkin Reed-­
Sternberg cells is an independent prognostic factor in classical Hodgkin’s lymphoma. J Clin Oncol
25:3101–3108
96. Nijland M, Veenstra RN, Visser L et al (2017) HLA
dependent immune escape mechanisms in B-cell
lymphomas: implications for immune checkpoint
inhibitor therapy? Oncoimmunology e1295202:6
97. Reichel J, Chadburn A, Rubinstein PG et al (2015)
Flow sorting and exome sequencing reveal the
oncogenome of primary Hodgkin and Reed-­
Sternberg cells. Blood 125:1061–1072
98. Diepstra A, Poppema S, Boot M et al (2008) HLA-G
protein expression as a potential immune escape
mechanism in classical Hodgkin’s lymphoma.
Tissue Antigens 71:219–226
99. Zocchi MR, Catellani S, Canevali P et al (2012)
High ERp5/ADAM10 expression in lymph node
microenvironment and impaired NKG2D ligands
recognition in Hodgkin lymphomas. Blood
119:1479–1489
100. Steidl C, Shah SP, Woolcock BW et al (2011)
MHC class II transactivator CIITA is a recurrent
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
113.
114.
85
gene fusion partner in lymphoid cancers. Nature
471:377–383
Walunas TL, Lenschow DJ, Bakker CY et al (1994)
CTLA-4 can function as a negative regulator of T
cell activation. Immunity 1:405–413
Alegre M, Frauwirth KA (2001) Thompson CB. T
cell regulation by CD28 and CTLA-4. Nat Rev
Immunol 1:220–228
Bashey A, Medina B, Corringham S et al (2009)
CTLA-4 blockade with ipilimumab to treat relapse
of malignancy after allogeneic hematopoietic cell
transplantation. Blood 113:1581–1588
Davids MS, Kim HT, Bachireddy P et al (2016)
Ipilimumab for patients with relapse after allogeneic
transplantation. N Engl J Med 375:143–153
Yamamoto R, Nishikori M, Kitawaki T et al (2008)
PD-1-PD-1 ligand interaction contributes to immunosuppressive microenvironment of Hodgkin lymphoma. Blood 15(111):3220–3224
Vari F, Arpon D, Keane C et al (2018) Immune evasion via PD-1/PD-L1 on NK-cells and monocytes/
macrophages is more prominent in Hodgkin lymphoma than DLBCL. Blood 131:1809–1819
Green MR, Monti S, Rodig SJ et al (2010) Integrative
analysis reveals selective 9p24.1 amplification,
increased PD-1 ligand expression, and further induction via JAK2 in nodular sclerosing Hodgkin lymphoma and primary mediastinal large B-cell. Blood
116:3268–3277
Ansell SM, Lesokhin AM, Borrello I et al (2015)
PD-1 blockade with nivolumab in relapsed or
refractory Hodgkin’s lymphoma. N Engl J Med
372:311–319
Armand P, Engert A, Younes A et al (2018)
Nivolumab for relapsed/refractory classical Hodgkin
lymphoma after failure of autologous hematopoietic
cell transplantation: extended follow-up of the multicohort single-arm phase II CheckMate 205 trial. J
Clin Oncol 36:1428–1439
Younes A, Santoro A, Shipp M et al (2016)
Nivolumab for classical Hodgkin’s lymphoma after
failure of both autologous stem-cell transplantation and brentuximab vedotin: a multicentre, multicohort, single-arm phase 2 trial. Lancet Oncol
17:1283–1294
Chen R, Zinzani PL, Fanale MA et al (2017) Phase
II study of the efficacy and safety of pembrolizumab
for relapsed/refractory classic Hodgkin lymphoma. J
Clin Oncol 35:2125–2132
Tumeh PC, Harview CL, Yearley JH et al (2014)
PD-1 blockade induces responses by inhibiting
adaptive immune resistance. Nature 515:568–571
Carey CD, Gusenleitner D, Lipschitz M et al (2017)
Topological analysis reveals a PD-L1-associated
microenvironmental niche for Reed-Sternberg cells
in Hodgkin lymphoma. Blood 130:2420–2430
Roemer MGM, Redd RA, Cader FZ et al (2018)
Major histocompatibility complex class II and
programmed death ligand 1 expression predict outcome after programmed death blockade
L. Visser et al.
86
115.
116.
117.
118.
119.
120.
121.
122.
123.
124.
125.
126.
127.
128.
in classical Hodgkin lymphoma. J Clin Oncol
36:942–950
Muenst S, Hoeller S, Dirnhofer S et al (2009)
Increased programmed death-1+ tumor-infiltrating
lymphocytes in classical Hodgkin lymphoma substantiate reduced overall survival. Hum Pathol
40:1715–1722
Goldberg MV, Drake CG (2011) LAG-3 in cancer immunotherapy. Curr Top Microbiol Immunol
344:269–278
Huard B, Prigent P, Tournier M et al (1995) CD4/
major histocompatibility complex class II interaction analyzed with CD4− and lymphocyte activation
gene-3 (LAG-3)-Ig fusion proteins. Eur J Immunol
25:2718–2721
Gandhi MK, Lambley E, Duraiswamy J et al (2006)
Expression of LAG-3 by tumor-infiltrating lymphocytes is coincident with the suppression of latent
membrane antigen-specific CD8+ T-cell function in
Hodgkin lymphoma patients. Blood 108:2280–2289
Camisaschi C, Casati C, Rini F et al (2010) LAG-3
expression defines a subset of CD4(+)CD25(high)
Foxp3(+) regulatory T cells that are expanded at
tumor sites. J Immunol 184:6545–6551
Duffield AS, Ascierto ML, Anders RA et al (2017)
Th17 immune microenvironment in Epstein-Barr
virus-negative Hodgkin lymphoma: implications for
immunotherapy. Blood Adv 1:1324–1334
Wein F, Weniger MA, Höing B et al (2017) Complex
immune evasion strategies in classical Hodgkin lymphoma. Cancer Immunol Res 5:1122–1132
Dukers DF, Jaspars LH, Vos W et al (2000)
Quantitative immunohistochemical analysis of
cytokine profiles in Epstein-Barr virus-positive
and -negative cases of Hodgkin’s disease. J Pathol
190:143–149
Herbst H, Foss HD, Samol J et al (1996) Frequent
expression of interleukin-10 by Epstein–Barr virus-­
harboring tumor cells of Hodgkin’s disease. Blood
87:2918–2929
Newcom SR, Kadin ME, Ansari AA et al (1988)
L-428 nodular sclerosing Hodgkin’s cell secretes
a unique transforming growth factor-beta active at
physiologic pH. J Clin Invest 82:1915–1921
Newcom SR, Tagra KK (1992) High molecular
weight transforming growth factor b is excreted in
the urine in active nodular sclerosing Hodgkin’s disease. Cancer Res 52:6768–6773
Marshall NA, Christie LE, Munro LR et al (2004)
Immunosuppressive regulatory T cells are abundant
in the reactive lymphocytes of Hodgkin lymphoma.
Blood 103:1755–1762
Juszczynski P, Ouyang J, Monti S et al (2007) The
AP1-dependent secretion of galectin-1 by Reed-­
Sternberg cells fosters immune privilege in classical Hodgkin lymphoma. Proc Natl Acad Sci U S A
104:13134–13139
Gandhi MK, Moll G, Smith C et al (2015) Brief
report Galectin-1 mediated suppression of Epstein-­
Barr virus—specific T-cell immunity in classic
Hodgkin lymphoma. Blood 110:1326–1330
129. Maggio EM, van den Berg A, de Jong D et al (2003)
Low frequency of FAS mutations in Reed-Sternberg
cells of Hodgkin’s lymphoma. Am J Pathol
162:29–35
130. Choe J-Y, Yun JY, Jeon YK et al (2014) Indoleamine
2,3-dioxygenase (IDO) is frequently expressed in
stromal cells of Hodgkin lymphoma and is associated with adverse clinical features: a retrospective
cohort study. BMC Cancer 14:335
131. Soliman H, Mediavilla-Varela M, Antonia S (2010)
Indoleamine 2,3-dioxygenase. Is i tan immune suppressor? Cancer J 16:354–359
132. Schwaller J, Tobler A, Niklaus G et al (1995)
Interleukin-12 expression in human lymphomas
and nonneoplastic lymphoid disorders. Blood
85:2182–2188
133. Niedobitek G, Pazolt D, Teichmann M et al (2002)
Frequent expression of the Epstein-Barr virus
(EBV)-induced gene, EBI3, an IL-12 p40-related
cytokine, in Hodgkin and Reed-Sternberg cells. J
Pathol 198:310–316
134. Steidl C, Lee T, Shah SP et al (2010) Tumor-­
associated macrophages and survival in classical
Hodgkin’s lymphoma. N Engl J Med 362:875–885
135. Guo B, Cen H, Tan X, Ke Q (2016) Meta-analysis of
the prognostic and clinical value of tumor-associated
macrophages in adult classical Hodgkin lymphoma.
BMC Med 14:159
136. Barros MHM, Segges P, Vera-Lozada G, Hassan
R, Niedobitek G (2015) Macrophage polarization
reflects T cell composition of tumor microenvironment in pediatric classical Hodgkin lymphoma and
has impact on survival. PLoS One 10:1–19
137. Hollander P, Rostgaard K, Smedby KE et al (2017)
An anergic immune signature in the tumor microenvironment of classical Hodgkin lymphoma is
associated with inferior outcome. Eur J Haematol
100:88–97
138. Andersen MD, Kamper P, Nielsen PS et al (2016)
Tumour-associated mast cells in classical Hodgkin’s
lymphoma: correlation with histological subtype,
other tumour-infiltrating inflammatory cell subsets
and outcome. Eur J Haematol 96:252–259
139. Alvaro T, Lejeune M, Salvado MT et al (2005)
Outcome in Hodgkin’s lymphoma can be predicted
from the presence of accompanying cytotoxic and
regulatory T cells. Clin Cancer Res 11:1467–1473
140. Kelley TW, Pohlman B, Elson P et al (2007) The
ratio of Foxp3+ regulatory T cells to Granzyme B+
cytotoxic T/NK cells predicts prognosis in classical
Hodgkin lymphoma and is independent of bcl-2 and
MAL expression. Am J Clin Pathol 128:958–965
141. Oudejans JJ, Jiwa NM, Kummer JA et al (1997)
Activated cytotoxic T cells as prognostic marker in
Hodgkin’s disease. Blood 89:1376–1382
142. Alonso-Álvarez S, Vidriales MB, Caballero MD
et al (2017) The number of tumor infiltrating
T-cell subsets in lymph nodes from patients with
Hodgkin lymphoma is associated with the outcome
after first line ABVD therapy. Leuk Lymphoma
58:1144–1152
5
What Have We Learnt
from Genomics
and Transcriptomics in Classic
Hodgkin Lymphoma
Davide Rossi and Christian Steidl
Contents
5.1
Introduction 87
5.2
5.2.1
5.2.2
5.2.3
5.2.4
enomics of Hodgkin and Reed-­Sternberg Cells G
Cytokine Signaling NF-κB Signaling PI3K/AKT/mTOR Signaling Immune Escape 88
88
89
89
90
5.3
The Transcriptome of HRS Cells 90
5.4
Microenvironment Profiling 91
5.5
Biomarker-Driven Prognostication
and Risk Stratification in cHL 92
Conclusions and Future Perspective 94
5.6
References 5.1
Introduction
A prominent pathological feature of cHL is the
abnormal immune response represented by the
abundant TME. It is thought that the majority of the
immune cells in the TME are recruited by a variety
of cytokines expressed by the HRS cells [1].
D. Rossi (*)
Oncology Institute of Southern Switzerland,
Bellinzona, Switzerland
e-mail: davide.rossi@ior.usi.ch
C. Steidl
Department of Pathology and Laboratory Medicine,
British Columbia Cancer Agency and the BC Cancer
Research Centre, Vancouver, BC, Canada
e-mail: CSteidl@bccancer.bc.ca
94
Cytokines are low-molecular-weight proteins with
a wide variety of functions that work either in a
paracrine manner to modulate the activity of surrounding cells or in an autocrine fashion to affect
the cells that produce them. Furthermore, it is a
widely accepted concept that the overexpression of
regulatory cytokines and TGFβ leads to a microenvironment that suppresses cell-mediated immunity
and in return favors HRS cell survival highlighting
the bidirectional crosstalk of cells involved in the
pathogenesis of HL [2].
The recent advances in HRS cell genomics
and profiling the tumor microenvironment have
already led to better insight into the molecular
underpinnings of the disease, and we are anticipating discovery of additional clues explaining
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_5
87
D. Rossi and C. Steidl
88
N=80
N=34
N=10
Circulating tumor DNA
(Spina et al.)
Laser microdissection
of Hodgkin and ReedSternberg cells
(Tiacci et al.)
Flow sorting of Hodgkin and
Reed-Sternberg cells cells
(Reichel et al.)
samples
47.1
32.4
5.9
11.8
23.5
20.6
5.9
0
2.9
8.8
17.6
0
0
0
2.9
0
0
0
2.9
11.8
0
0
2.9
0
0
5.9
0
0
0
0
0
0
0
0
0
0
2.9
0
2.9
0
0
0
0
0
0
2.9
0
0
0
0
0
0
2.9
0
0
0
0
2.9
0
0
0
0
0
2.9
0
0
5.9
0
0
0
0
0
0
2.9
0
0
0
0
40
10
60
40
20
70
0
10
0
0
10
10
50
0
0
0
0
10
0
10
0
0
10
10
0
0
0
0
10
0
0
0
0
0
0
0
0
0
0
0
0
20
0
0
0
0
0
0
0
0
10
0
10
0
10
0
0
10
0
0
10
0
0
10
0
0
0
10
0
0
0
0
0
0
10
0
0
0
% of mutated samples
NA
37.5
35.0
27.5
18.8
16.3
15.0
12.5
11.3
11.3
11.3
10
10
10
7.5
7.5
7.5
6.3
6.3
6.3
6.3
5
5
5
5
5
5
5
3.8
3.8
3.8
3.8
3.8
3.8
3.8
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
2.5
1.3
1.3
1.3
1.3
1.3
1.3
1.3
1.3
0
0
0
0
0
0
0
0
0
0
0
SOCS1
STAT6
TNFAIP3
ITPKB
GNA13
B2M
ATM
SPEN
KMT2D
TP53
XPO1
BTG1
HIST1H1E
ARID1A
CD36
FBXW7
P2RY8
CIITA
IKBKB
NFKBIE
PIK3CA
BIRC3
CREBBP
FBXO11
IRF8
TBL1XR1
PCBP1
TNFRSF14
BTK
CD58
HIST1H1C
RIPK1
TRAF3
ZMYM3
MYC
BCL2
BCOR
CCND3
CD79B
CXCR4
EGR2
HIST1H3B
ID3
IRAK1
IRF4
KLHL6
MAP2K1
MAP3K14
NOTCH1
NOTCH2
PIM1
PLCG2
POT1
S1PR2
SIN3A
TET2
WHSC1
EZH2
PRDM1
EP300
BCL6
BRAF
CCND1
FGFR2
KLF2
MEF2B
STAT3
CARD11
CCND2
CD70
CD79A
FOXO1
KRAS
MYD88
NRAS
RRAGC
TCF3
TRAF2
Spina et al. cHL cohort
Tiacci et al. cHL cohort
Reichel et al. cHL cohort
Not Available
Mutated sample
Unmutated sample
Fig. 5.1 The mutational profile of newly diagnosed
cHL. The heatmap shows individual non-synonymous
somatic mutations detected in three different cohorts
(Spine et al., green; Tiacci et al., yellow; Reichel et al.,
blue). Each cohort has a different source of tumor DNA
(i.e., circulating tumor DNA, DNA from laser microdissected Hodgkin and Reed-Sternberg cells, and DNA from
flow-sorted Hodgkin and Reed-Sternberg cells). Each row
represents a gene and each column represents a primary
tumor. The heatmap was manually clustered to emphasize
mutational co-occurrence. Mutations are color-coded in
red. The horizontal bar graph shows the gene mutation
frequency found in each different cohort
the unique crosstalk and symbiosis of the malignant cells with the non-malignant cells in the
TME. In the following, we will highlight recent
advances and future directions in (1) HRS cell
genomics (Fig. 5.1) and (2) gene expression
profiling.
evidence that various molecular mechanisms,
including gene mutations and chromosomal
alterations, can converge along with deregulated
surface receptor signaling to lead to exuberant
activation of the Janus kinase-signal transducer
and activator of transcription (JAK-STAT) pathway [3–5].
Chromosomal aberrations of the JAK2 locus
on 9p24.1 in HRS cells were reported in one
study in the large majority of cHL cases, including copy gain in 60% of cases, amplification in
30%, and polysomy in 10% [5]. Almost ubiquitous (∼90% of cases) are genetic alterations of a
variety of other JAK-STAT pathway members,
which goes beyond previous estimates based on
the presence of copy number gains of JAK2.
5.2
enomics of Hodgkin and
G
Reed-­Sternberg Cells
5.2.1
Cytokine Signaling
Constitutive activation of cytokine signaling
pathways is a long recognized molecular hallmark of HRS cells. A number of studies provided
5
What Have We Learnt from Genomics and Transcriptomics in Classic Hodgkin Lymphoma
These include mutational disruption of the
SOCS1 (40%) and PTPN1 (20%) negative pathway regulators, activating mutations of JAK1
(10%), and multiple STAT transcription factors
(STAT6, 30%; STAT3, 10% STAT5B, 10%) [4].
The association between convergent and
recurrent point mutations in genes coding for
interacting proteins of the JAK-STAT pathway
is a common mechanism shared by CD30+
lymphomas, in particular cHL and anaplastic
large cell lymphoma. Concurrence of these
multiple somatic events indicates that these
synergistic mutations are strongly selected for
beyond single alterations to sustain pathway
activation [6].
The pervasive targeting of JAK-STAT signaling genes in cHL, along with functional genomic
studies, confirmed that JAK-STAT pathway activation represents a vulnerability of cHL and
makes clinically available JAK or STAT inhibitors an attractive therapeutic approach in this disease [4].
5.2.2
NF-κB Signaling
Overall, genetic lesions in the NF-κB pathway
occur in most of cHL cases, confirming their
important role in the pathogenesis of this disease. Genomic gains/amplifications of the
NF-κB transcription factor REL have been
described in about 70% of cHL cases causing
protein overexpression [7].
Mutations in negative regulators of NF-κB
constitute a second important mechanism of
pathway activation. NFKBIA, encoding IκBα, an
inhibitor that binds NF-κB factors and prevents
their nuclear translocation, is mutated in about
20% of cHL [8]. NFKBIE, encoding IκBε, an
inhibitor that binds NF-κB factors and prevents
their nuclear translocation, has been found in
30% of cases [9]. TNFAIP3 the master negative
regulator of NF-κB pathway is mutated in 30% of
cases [3, 10].
Overall, NF-κB pathway mutations have
been described in cHL with a higher frequency
in EBV-negative cases, consistent with data
establishing expression of the EBV-latent
89
membrane protein 1 (LMP-1) as an independent contributor to constitutive activation of
NF-κB in cHL [11, 12].
5.2.3
PI3K/AKT/mTOR Signaling
Mutations within the PI3K/AKT/mammalian target of rapamycin (mTOR) pathway occur in 50%
of cHL, consistent with the pre-clinical evidence
that cHL is addicted to this actionable cellular
program [13]. ITPKB is mutated in 25% of cases.
ITPKB is a non-canonical antagonist of
PI3K. Physiologically, ITPKB dampens PI3K/
AKT signaling by producing IP4, a soluble
antagonist of the AKT-activating PI3K-product
PIP3.
ITPKB mutations are quite specific for cHL,
being rare or absent in other lymphomas, and
cause the subcellular delocalization of the
mutated protein in primary HRS cells. Moreover,
ITPKB mutations correlate with PI3K/AKT signaling activation at both the gene expression and
protein levels and, consistent with linkages to the
downstream PI3K pathway, associate with resistance to PI3K inhibitors [3, 4].
The Ga13 G-protein subunit encoded by
GNA13 is mutated in 10% of cHL [3, 4]. By
transmitting signals from the G-proteincoupled receptors S1PR2 and P2RY8 that
result in the inhibition of AKT phosphorylation, Ga13 ensures the proper confinement of
proliferating germinal center (CG) B cells
within secondary lymphoid follicles and at the
same time constrains their expansion by facilitating apoptosis in this potentially dangerous
niche. Inactivating GNA13 mutations promote
altered GC B-cell migration within and beyond
the GC, as well as impaired cellular adhesion,
resulting in cells that may have a reduced ability to establish interactions with GC helper
cells. Under normal conditions, a GC cell that
is unable to form these helper cell interactions,
due to either GC exit or ineffective cellular
adhesion, would undergo apoptosis. However,
GNA13-mutated GC B cells are resistant to
programmed cell death by leading to elevated
levels of pAKT [14].
D. Rossi and C. Steidl
90
Importantly, the genomic studies of microdissected HRS cells and ctDNA strongly suggest
that mutations of STAT6, TNFAIP3, GNA13, and
ITPKB are preferentially occurring in the ancestral clones, indicating that they are an early event
in cHL pathogenesis [3, 4].
5.3
The Transcriptome
of HRS Cells
Overall, gene expression profiling experiments
have contributed substantially to an improved
understanding of the disease with respect to the
inherent phenotypic features of the malignant
HRS cells and the specific composition of the
5.2.4 Immune Escape
tumor microenvironment. Furthermore, first
steps could be made to establish outcome corClassical HL leverages multiple genetic mecha- relations with the potential to improve treatment
nisms to escape immunosurveillance. First, reduc- outcome prediction. However, many questions
tion or loss of antigen presentation through B2M remain including often contradictory results
inactivating mutations/deletion has been described derived from different patient cohorts. Focusing
in 30% of cases [4, 15]. B2M encodes β2 micro- on HRS cells, the first major contribution of
globulin, a key component of the major histocom- gene expression profiling was made by investipatibility complex (MHC) class I which is gating HL-derived cell lines. These pivotal studrequired for its expression and antigen presenta- ies first established a transcriptome-wide view
tion on the cell surface. Consistently, genetic dis- of the malignant cell compartment describing a
ruption of B2M results in the loss of MHC class I unifying gene signature for cHL [22]. Together
protein expression on lymphoma cells [16, 17].
with other important similar studies, this gene
Second, gene rearrangements involving the expression work helped to elucidate the loss of
MHC class II transactivator CIITA were found in B-cell signature phenotypes and the deregulated
15% of cases. CIITA rearrangements result in the expression of transcription factor networks in
disruption of its transcriptional proprieties and comparison to the normal germinal center B-cell
loss of MHC class II expression on cHL cells. counterparts [23–26]. Major advances have also
Both MHC class I and MHC class II losses are been made examining microdissected HRS cells
predicted to abrogate the interaction of the T-cell from clinical biopsy material that further charreceptor (TCR) with a MHC-bound antigen pre- acterized transcriptional changes in primary
sented on the cell surface, which is the first signal cells [27–29]. Steidl and colleagues identified
required to activate T-cell antitumor response [18]. significant phenotypic heterogeneity within
Loss of both MHC I and II expression and related cHL and described for the first time genomelack of neoantigen expression have been consis- wide association with treatment outcome [28]
tently found to induce “cold” immune microenvi- (Fig. 5.2). The second study by Tiacci and colronments in lymphoma and other cancers [19, 20]. leagues added significant texture to the primary
Third, PD-L1 and PD-L2 overexpression HRS cell expression phenotype emphasizing the
driven by copy gain of 9p24.1 is a frequent event differences in comparison to HL-derived cell
in cHL. Alterations of the PD-L1 and PD-L2 loci lines [29]. Furthermore, two molecularly diswere reported to include polysomy in 5% of cHL, tinct cHL subtypes were discovered related to
copy gain in 56%, and amplification in 36%. The the transcription factor activity of NOTCH1,
9p24.1 amplification in cHL acts through two MYC, and IRF4. Another study for the first time
distinct mechanisms resulting in copy number-­ also focused on gene expression profiling of
dependent increases of PD-L1 and PD-L2 expres- microdissected cells from nodular lymphocyte
sion and increased JAK/STAT signaling promoted predominant Hodgkin lymphoma (NLPHL)
by JAK2 protein expression which is almost describing a close relationship to classical
exclusively co-regulated with PD-L1 and Hodgkin lymphoma and T-cell-­rich B-cell lymPD-L2 in the 9p24.1 amplicon [21].
phoma [27].
5
What Have We Learnt from Genomics and Transcriptomics in Classic Hodgkin Lymphoma
a
91
b
Cluster
HRS Cell Status
Histological Subtype
Sample Type
3
Treatment Outcome
Histological Subtype
HRS Cell Status
Sample Type
4
2
2
1
0
0
-1
-2
-2
Residual B-cell
Signature
Receptor Signaling Signature
Cluster
-4
-8
IGL
AMIGO2
IGJ
IGLJ3
HLA-C
IGHG3
-12
POU2AF1
PERP
RANK
0
TGFBR3
IL9
IL26
CCND2
TFPI2
Fold Change
200
150
100
50
0
CCL17
Fold Change
C1QB
GZMB
CXCL10
Fold Change
Cytotoxic
Signature
20
15
10
5
0
HRS Cell Status
Germinal Centre
EBV Positive
Cluster A
EBV Negative
Cluster B
Not determined
Cluster C
Histological Subtype
Sample Type
Nodular Sclerosis
Diagnosis
Mixed Cellularity
Relapse
Treatment Outcome
Treatment Success
Treatment Failure
GrB Immunohistochemistry
RANK (TNFRSF11A) Immunohistochemistry
Fig. 5.2 Expression profiling of 29 samples of microdissected Hodgkin and Reed-Sternberg cells. (a) Unsupervised
hierarchical clustering of gene expression profiles is
shown using high variance genes. Red indicates relative
overexpression and green relative under-expression.
Patient clusters, histological subtype, EBV positivity of
HRS cells by EBER in situ hybridization, and sample type
are shown. The average fold changes of genes representative of the three main signatures are shown in the bar
plots. Representative immunohistochemistry images are
depicted demonstrating cytoplasmic positivity of
Granzyme B (GrB, black arrows) and RANK in HRS
cells. (b) Unsupervised hierarchical clustering of the
cohort using the most differentially expressed genes
between primary treatment failure and success. Treatment
outcome, histological subtype, EBV positivity of HRS
cells by EBER in situ hybridization, and sample type are
shown. Cases cluster according to the outcome groups
(two main clusters)
5.4
ment [30–33]. However, some of these data provide evidence that at least parts of the apparent
signatures are derived from HRS cells [31, 33]. In
one study a specific gene expression signature
could be linked to EBV positivity with genes
overexpressed indicative of an increased Th1/
antiviral response in comparison to the EBV-­
negative cases [32]. In addition to a better characterization of certain Hodgkin lymphoma subtypes
Microenvironment Profiling
Focusing on the HL microenvironment, a number
of genome-wide gene expression studies have
been published to date analyzing whole tissue
lymph node biopsy material. Since the HRS cells
are largely outnumbered by reactive cells in most
biopsies, these studies on whole frozen biopsies
are regarded as a reflection of the microenviron-
D. Rossi and C. Steidl
92
two chemotherapy courses has prognostic implications. A drop of 100-fold or 2-log drop in
ctDNA after two chemotherapy courses, a
threshold proposed and validated also in
DLBCL, associates with complete response and
5.5
Biomarker-Driven
cure in advanced-stage cHL treated with ABVD
[3]. Conversely, a drop of less than 2-log in
Prognostication and Risk
ctDNA after two ABVD courses associates with
Stratification in cHL
progression and inferior survival. Quantification
The lack of extensive genotyping of microdis- of ctDNA complements interim PET/CT in
sected HRS cells from large cHL patient cohorts determining residual disease. Indeed, cured
has so far limited the identification of mutations patients who are inconsistently judged as
affecting cHL outcome. ctDNA has been estab- interim PET/CT positive have a >2-log drop in
lished as a source of tumor DNA for cHL muta- ctDNA, while relapsing patients who are incontional profiling. By overcoming the major sistently judged as interim PET/CT negative
technical hurdles that have so far limited cHL have a <2-log drop in ctDNA. On this basis,
genotyping, ctDNA technology will allow large-­ incorporation of both PET/CT and ctDNA monscale assessment of mutations in different clini- itoring into clinical trials should allow to precal phases ranging from newly diagnosed to cisely define their cumulative sensitivity and
refractory disease, and longitudinally during dis- specificity in anticipating the clinical course of
ease treatment, which in turn can reveal yet cHL patients. Indeed, though interim PET/CT
unknown prognostic and predictive biomarkers response assessment is a novel approach to
for cHL [3] (Fig. 5.3).
refine management strategies before completing
Beside disclosing tumor mutation profiles, treatment in cHL, meta-analyses demonstrated a
ctDNA can also provide an estimate of the lev- certain degree of inaccuracy of this application.
els of residual disease during treatment in In order to fill this gap, an area of growing intercHL. Consistently, ctDNA quantification after est is pairing interim PET/CT with biomarkers,
Log fold change in tumor ctDNA
defined by specific gene signatures, these experiments also allowed for the study of outcome correlations using supervised analyses.
1
0
Nodular sclerosis cHL (NSCHL)
Mixed cellularity cHL (MCCHL)
Lymphocyte rich cHL (LRCHL)
-1
-2
-3
iPET positive – Progressive disease
iPET positive – Cured
iPET negative – Progressive disease
iPET negative – Cured
-4
ND
3
4
4
3
3
3
3
1
2
2
2
4
4
5
4
1
2
1
1
1
1
1
3
1
PD PD PD PD PD PD CR CR CR CR CR CR CR CR CR CR CR CR CR CR CR CR CR CR
Fig. 5.3 Change in tumor ctDNA is a prognostic biomarker in cHL treated with chemotherapy. Waterfall plot
of the log-fold change in ctDNA load after two courses of
ABVD in 24 advanced-stage cHL cases. At the bottom of
the graph, the interim PET/CT response scored according
to the Deauville criteria, and the final outcome of the
patient is indicated. Histological subtype of cHL is shown
Deauville score
Outcome
above the plot. Each column is color-coded according to
the interim PET/CT results and the final patient outcome.
Levels of ctDNA are normalized to baseline levels. The
dashed line tracks the −2-log threshold (iPET interim
PET/CT, ND not detectable, PD progressive disease, CR
complete remission and cure)
5
What Have We Learnt from Genomics and Transcriptomics in Classic Hodgkin Lymphoma
such as ctDNA or serum TARC, to enhance their
cumulative predictive value.
The type of 9p24.1 chromosomal aberration
affects cHL outcome in both chemotherapy and
immunotherapy treatment settings. Among
chemotherapy-­
treated cHL, 9p24.1 amplification, but not polysomy or copy gain, associates
with inferior progression-free survival [21].
Among patients treated with checkpoint blockade antibodies, those with higher-level 9p24.1
alterations and PD-L1 expression on HRS cells
had superior PFS [34]. These analyses highlight
the importance of quantifying and specifically
delineating PD-L1 expression in malignant HRS
cells for prognostic purposes.
Beside genetics, the tumor/TME phenotype
has been prominently involved in past and ongoing biomarker considerations in cHL. Studies
have used dichotomized clinical data sets based
on slightly different definitions of clinical
extremes according to the outcome after systemic
treatment (i.e., treatment success versus treatment failure). However, these types of analyses
have in part yielded conflicting results regarding
the specific signatures that best define these clinical extremes. While one study found overexpression of genes involved in fibroblast activation,
angiogenesis, extracellular matrix remodeling,
and downregulation of tumor suppressor genes to
be linked with an unfavorable prognosis, another
study found a correlation of fibroblast activation,
fibroblast chemotaxis, and matrix remodeling
with improved outcome [30, 31]. While small
sample sizes in both studies might have hampered interpretation, a more recent study investigated gene expression profiles of 130 patients
including 38 patients whose primary treatments
failed [33]. This study validated previously
reported outcome correlations and furthermore
showed that a gene signature of macrophages
was linked to primary treatment failure. In a
number of immunohistochemistry-based follow­up studies, multiple groups demonstrated that the
enumeration of CD68+ macrophages in lymph
node biopsies was a strong and independent predictor of disease-specific survival [35].
Specifically, an elegant retrospective study using
Intergroup E2496 trial material (comparing
93
ABVD to the Stanford V regimen) showed that
high abundance of both CD68+ and CD163+
cells was correlated with shorter progression-free
and overall survival independent of the IPS [36].
Importantly, the latter study used a computer-­
based scoring algorithm (Aperio) and systematically derived scoring thresholds that were tested
in an independent validation cohort. Maximizing
the concept of combining markers for building
outcome predictors, a recent study used the same
E2496 trial material to train a predictive model
using intermediate density digital gene expression profiling developed in and applicable to routinely collected formalin-fixed paraffin-embedded
tissue [37]. In this study the authors developed a
23-gene predictive model and associated thresholds to distinguish high-risk from low-risk
advanced-stage Hodgkin lymphoma using overall survival as the end point. Encouragingly,
when applied to an independent cohort treated
with ABVD chemotherapy, the model validated
the results in the E2496 training cohort identifying the patient at high risk of death. Follow-up
studies are needed to further validate and implement biomarker assays for potential routine clinical use, risk stratification, and assessment as a
predictive biomarker possibly guiding initial
treatment decisions.
To date, cHL research has been for the most
part focused on primary specimens, and only a
few studies have explored the biology of relapse.
However recently, the feasibility of biomarker
studies and assay development at the time point
of relapse was demonstrated in the context of
outcome prediction of salvage therapy and
ASCT [38]. The authors demonstrated that gene
expression patterns, reflecting TME composition, differ significantly between matched primary and relapse specimens in a subset of cHL
patients. Based on the superior predictive properties of gene expression measurements in
relapse specimens, a novel clinically applicable
prognostic model/assay (RHL30) was developed that identifies a subset of patients at high
risk of treatment failure following salvage therapy and ASCT. Specifically, RHL30 identifies a
high-risk group of patients with significantly
inferior post-­ASCT-­FFS compared to the low-
D. Rossi and C. Steidl
94
risk group (5-year: 23.8% high-risk vs. 77.5%
low-risk) and also inferior post-ASCT-OS
(5-year: 28.7% high-­risk vs. 85.4% low-risk).
Importantly, the prognostic power of RHL30
was reproduced in two separate validation
cohorts of relapse specimens, and the RHL30
was statistically independent of all previously
described prognostic markers in the validation
cohorts, including post-salvage therapy response
assessment by PET/CT [38].
5.6
Conclusions and Future
Perspective
The advent of next-generation sequencing has
significantly added to the armamentarium of
genomics techniques interrogating tumor genetics of cHL and elucidating the molecular underpinnings of the unique crosstalk of the malignant
HRS cells with their immune microenvironment.
The sequencing studies of ctDNA and enrichment of HRS cells confirmed the importance of,
and added texture to, the known molecular hallmarks of NFκB, JAK-STAT, and PI3K signaling
as well as immune privilege phenotypes.
Moreover, gene expression profiling studies of
the microenvironment have reached more maturity in comprehensively describing cellular compartments in the TME and validated key
correlations to pathologic and clinical outcome
data. In particular, effective biomarker assay
translation appears more and more realistic with
the emergence of methods that are compatible
with FFPE tissues that can be applied to relapse
biopsies and are minimally invasive (e.g., serial
peripheral blood draws) for dynamic biomarker
testing. Despite these most recent advances, a
number of challenges and open questions remain
that need to be addressed in future studies. First,
with respect to cHL biology, no unique and specific somatic gene mutations have been identified that would explain the unique histopathology
of cHL in contrast to other lymphomas, leaving
room for future discoveries. Second, systematic
integration of HRS cell genomics with features
and cellular components of the TME are lacking.
Third, sample numbers for genomic landscape
studies are still limited to be fully powered for
mutational pattern analysis and robust outcome
correlates in patients treated with standard of
care. Finally, with the emergence of targeted
therapies (e.g., brentuximab vedotin [39]) and
modern immunotherapies (e.g., checkpoint
inhibitors [40] or bispecific antibodies [41]),
predictive biomarker development using genomics has to be prioritized alongside the next generation of clinical trials and population-based
outcome studies of patients receiving these novel
therapies in the standard of care setting.
Excitingly, novel cutting-edge genomics techniques might also overcome some of the
described obstacles, including HRS cell sequencing, to interrogate the non-coding space (e.g.,
whole genome sequencing), epigenetic profiling
(e.g., ATAC-seq, bisulfite sequencing), and
RNAseq at the single cell level to characterize
the TME. Integrating these novel genomics
approaches for dynamic, multi-time point biomarker testing alongside existing and novel therapeutic approaches holds the great promise to
fully realize the benefits of precision medicine
by genomics-driven clinical decision-making.
References
1. Kuppers R (2009) The biology of Hodgkin's lymphoma. Nat Rev Cancer 9(1):15–27
2. Steidl C, Connors JM, Gascoyne RD (2011) Molecular
pathogenesis of Hodgkin’s lymphoma: increasing evidence of the importance of the microenvironment. J
Clin Oncol 29(14):1812–1826
3. Spina V, Bruscaggin A, Cuccaro A, Martini M, Di
Trani M, Forestieri G et al (2018) Circulating tumor
DNA reveals genetics, clonal evolution, and residual disease in classical Hodgkin lymphoma. Blood
131(22):2413–2425
4. Tiacci E, Ladewig E, Schiavoni G, Penson A, Fortini
E, Pettirossi V et al (2018) Pervasive mutations of
JAK-STAT pathway genes in classical Hodgkin lymphoma. Blood 131(22):2454–2465
5. Roemer MG, Advani RH, Redd RA, Pinkus GS,
Natkunam Y, Ligon AH et al (2016) Classical
Hodgkin lymphoma with reduced beta2M/MHC
class I expression is associated with inferior outcome independent of 9p24.1 status. Cancer Immunol
Res 4(11):910–916
6. Crescenzo R, Abate F, Lasorsa E, Tabbo F, Gaudiano
M, Chiesa N et al (2015) Convergent mutations and
kinase fusions lead to oncogenic STAT3 activation in anaplastic large cell lymphoma. Cancer Cell
27(4):516–532
5
What Have We Learnt from Genomics and Transcriptomics in Classic Hodgkin Lymphoma
7. Joos S, Menz CK, Wrobel G, Siebert R, Gesk S, Ohl
S et al (2002) Classical Hodgkin lymphoma is characterized by recurrent copy number gains of the short
arm of chromosome 2. Blood 99(4):1381–1387
8. Lake A, Shield LA, Cordano P, Chui DT, Osborne J,
Crae S et al (2009) Mutations of NFKBIA, encoding IkappaBalpha, are a recurrent finding in classical
Hodgkin lymphoma but are not a unifying feature of
non-EBV-associated cases. Int J Cancer 125:1334
9. Emmerich F, Theurich S, Hummel M, Haeffker A,
Vry MS, Dohner K et al (2003) Inactivating I kappa B
epsilon mutations in Hodgkin/Reed-Sternberg cells. J
Pathol 201(3):413–420
10. Schmitz R, Hansmann ML, Bohle V, Martin-Subero
JI, Hartmann S, Mechtersheimer G et al (2009)
TNFAIP3 (A20) is a tumor suppressor gene in
Hodgkin lymphoma and primary mediastinal B cell
lymphoma. J Exp Med 206(5):981–989
11. Schumacher MA, Schmitz R, Brune V, Tiacci E,
Doring C, Hansmann ML et al (2010) Mutations in
the genes coding for the NF-kappaB regulating factors IkappaBalpha and A20 are uncommon in nodular lymphocyte-predominant Hodgkin's lymphoma.
Haematologica 95(1):153–157
12. Etzel BM, Gerth M, Chen Y, Wunsche E, Facklam T,
Beck JF et al (2017) Mutation analysis of tumor necrosis factor alpha-induced protein 3 gene in Hodgkin
lymphoma. Pathol Res Pract 213(3):256–260
13. Johnston PB, Pinter-Brown LC, Warsi G, White K,
Ramchandren R (2018) Phase 2 study of everolimus
for relapsed or refractory classical Hodgkin lymphoma. Exp Hematol Oncol 7:12
14. Muppidi JR, Schmitz R, Green JA, Xiao W, Larsen
AB, Braun SE et al (2014) Loss of signalling via
Galpha13 in germinal centre B-cell-derived lymphoma. Nature 516(7530):254–258
15. Reichel J, Eng K, Elemento O, Cesarman E, Roshal M
(2013) Exome sequencing of purified Hodgkin Reed-­
Sternberg cells reveals recurrent somatic mutations in
genes responsible for antigen presentation, chromosome integrity, transcriptional regulation and protein
ubiquitination. Blood 122(21):625
16. Liu Y, Abdul Razak FR, Terpstra M, Chan FC, Saber
A, Nijland M et al (2014) The mutational landscape of
Hodgkin lymphoma cell lines determined by whole-­
exome sequencing. Leukemia 28(11):2248–2251
17. Challa-Malladi M, Lieu YK, Califano O, Holmes AB,
Bhagat G, Murty VV et al (2011) Combined genetic
inactivation of beta2-microglobulin and CD58 reveals
frequent escape from immune recognition in diffuse
large B cell lymphoma. Cancer Cell 20(6):728–740
18. Steidl C, Shah SP, Woolcock BW, Rui L, Kawahara
M, Farinha P et al (2011) MHC class II transactivator
CIITA is a recurrent gene fusion partner in lymphoid
cancers. Nature 471(7338):377–381
19. Ennishi D, Takata K, Beguelin W, Duns G, Mottok
A, Farinha P et al (2019) Molecular and genetic characterization of MHC deficiency identifies EZH2 as
therapeutic target for enhancing immune recognition.
Cancer Discov 9(4):546–563
95
20. Grasso CS, Giannakis M, Wells DK, Hamada T, Mu
XJ, Quist M et al (2018) Genetic mechanisms of
immune evasion in colorectal Cancer. Cancer Discov
8(6):730–749
21. Roemer MG, Advani RH, Ligon AH, Natkunam Y,
Redd RA, Homer H et al (2016) PD-L1 and PD-L2
genetic alterations define classical Hodgkin lymphoma and predict outcome. J Clin Oncol 34:2690
22. Kuppers R, Klein U, Schwering I, Distler V,
Brauninger A, Cattoretti G et al (2003) Identification
of Hodgkin and Reed-Sternberg cell-specific
genes by gene expression profiling. J Clin Invest
111(4):529–537
23. Schwering I, Brauninger A, Klein U, Jungnickel
B, Tinguely M, Diehl V et al (2003) Loss of the
B-lineage-specific gene expression program in
Hodgkin and Reed-Sternberg cells of Hodgkin lymphoma. Blood 101(4):1505–1512
24. Mathas S, Janz M, Hummel F, Hummel M, Wollert-­
Wulf B, Lusatis S et al (2006) Intrinsic inhibition of transcription factor E2A by HLH proteins
ABF-1 and Id2 mediates reprogramming of neoplastic B cells in Hodgkin lymphoma. Nat Immunol
7(2):207–215
25. Stein H, Marafioti T, Foss HD, Laumen H, Hummel
M, Anagnostopoulos I et al (2001) Down-regulation
of BOB.1/OBF.1 and Oct2 in classical Hodgkin disease but not in lymphocyte predominant Hodgkin
disease correlates with immunoglobulin transcription.
Blood 97(2):496–501
26. Jundt F, Kley K, Anagnostopoulos I, Schulze
Probsting K, Greiner A, Mathas S et al (2002) Loss
of PU.1 expression is associated with defective
immunoglobulin transcription in Hodgkin and Reed-­
Sternberg cells of classical Hodgkin disease. Blood
99(8):3060–3062
27. Brune V, Tiacci E, Pfeil I, Doring C, Eckerle S, van
Noesel CJ et al (2008) Origin and pathogenesis of
nodular lymphocyte-predominant Hodgkin lymphoma as revealed by global gene expression analysis. J Exp Med 205(10):2251–2268
28. Steidl C, Diepstra A, Lee T, Chan FC, Farinha P, Tan
K et al (2012) Gene expression profiling of microdissected Hodgkin Reed-Sternberg cells correlates with
treatment outcome in classical Hodgkin lymphoma.
Blood 120(17):3530–3540
29. Tiacci E, Doring C, Brune V, van Noesel CJ, Klapper
W, Mechtersheimer G et al (2012) Analyzing primary Hodgkin and Reed-Sternberg cells to capture
the molecular and cellular pathogenesis of classical
Hodgkin lymphoma. Blood 120(23):4609–4620
30. Devilard E, Bertucci F, Trempat P, Bouabdallah R,
Loriod B, Giaconia A et al (2002) Gene expression
profiling defines molecular subtypes of classical
Hodgkin’s disease. Oncogene 21(19):3095–3102
31. Sanchez-Aguilera A, Montalban C, de la Cueva P,
Sanchez-Verde L, Morente MM, Garcia-Cosio M
et al (2006) Tumor microenvironment and mitotic
checkpoint are key factors in the outcome of classic
Hodgkin lymphoma. Blood 108(2):662–668
96
32. Chetaille B, Bertucci F, Finetti P, Esterni B,
Stamatoullas A, Picquenot JM et al (2009) Molecular
profiling of classical Hodgkin lymphoma tissues
uncovers variations in the tumor microenvironment
and correlations with EBV infection and outcome.
Blood 113(12):2765–3775
33. Steidl C, Lee T, Shah SP, Farinha P, Han G, Nayar T
et al (2010) Tumor-associated macrophages and survival in classic Hodgkin's lymphoma. N Engl J Med
362(10):875–885
34. Roemer MGM, Redd RA, Cader FZ, Pak CJ,
Abdelrahman S, Ouyang J et al (2018) Major histocompatibility complex class II and programmed
death ligand 1 expression predict outcome after programmed death 1 blockade in classic Hodgkin lymphoma. J Clin Oncol 36(10):942–950
35. Steidl C, Farinha P, Gascoyne RD (2011) Macrophages
predict treatment outcome in Hodgkin’s lymphoma.
Haematologica 96(2):186–189
36. Tan KL, Scott DW, Hong F, Kahl BS, Fisher RI,
Bartlett NL et al (2012) Tumor-associated macrophages predict inferior outcomes in classic Hodgkin
lymphoma: a correlative study from the E2496 intergroup trial. Blood 120(16):3280–3287
D. Rossi and C. Steidl
37. Scott DW, Chan FC, Hong F, Rogic S, Tan KL,
Meissner B et al (2013) Gene expression-based model
using formalin-fixed paraffin-embedded biopsies
predicts overall survival in advanced-stage classical
hodgkin lymphoma. J Clin Oncol 31(6):692–700
38. Chan FC, Mottok A, Gerrie AS, Power M, Nijland
M, Diepstra A et al (2017) Prognostic model to predict post-autologous stem-cell transplantation outcomes in classical Hodgkin lymphoma. J Clin Oncol
35(32):3722–3733
39. Connors JM, Jurczak W, Straus DJ, Ansell SM, Kim
WS, Gallamini A et al (2018) Brentuximab Vedotin
with chemotherapy for stage III or IV Hodgkin’s lymphoma. N Engl J Med 378(4):331–344
40. Ansell SM, Lesokhin AM, Borrello I, Halwani A,
Scott EC, Gutierrez M et al (2015) PD-1 blockade
with nivolumab in relapsed or refractory Hodgkin’s
lymphoma. N Engl J Med 372(4):311–319
41. Rothe A, Sasse S, Topp MS, Eichenauer DA, Hummel
H, Reiners KS et al (2015) A phase 1 study of the
bispecific anti-CD30/CD16A antibody construct
AFM13 in patients with relapsed or refractory
Hodgkin lymphoma. Blood 125(26):4024–4031
Part II
Diagnosis and First-Line Treatment
6
Clinical Evaluation
James O. Armitage and Jonathan W. Friedberg
Contents
6.1
Presenting Manifestations 6.2
Physical Findings and Laboratory Abnormalities 101
6.3
Pathologic Diagnosis: The Biopsy 101
6.4
Staging Systems for Hodgkin Lymphoma 103
6.5
Imaging Evaluation of the Extent of Disease 106
6.6
Clinical Evaluation During Therapy 107
6.7
Definition of the Response to Treatment 107
6.8
Complete Remission 107
6.9
Follow-Up Management 108
6.10
Conclusion 108
References 6.1
Presenting Manifestations
Hodgkin lymphoma can come to clinical attention
in a variety of ways. These include symptoms
caused by a growing mass and systemic symptoms
that are presumably cytokine induced, and a diagnosis can be made incidentally as part of an evaluJ. O. Armitage (*)
Division of Hematology/Oncology, Nebraska
Medical Center, Omaha, NE, USA
e-mail: joarmita@unmc.edu
J. W. Friedberg
Wilmot Cancer Institute, University of Rochester,
Rochester, NY, USA
99
108
ation for an unrelated problem. By far the most
common presentation of Hodgkin lymphoma is the
enlargement of lymph nodes that is typically painless and progressive. Although the most common
place for lymph nodes to be found is in the neck
and supraclavicular region, any lymph node-bearing area can be involved. Patients typically find
enlarged nodes above the clavicle and seek medical
attention when they do not regress, while physicians are relatively more likely to discover lymph
nodes in other areas as part of a physical examination. Mediastinal lymphadenopathy is a particularly common finding in young women with
Hodgkin lymphoma. This might be found incidentally on a chest X-ray or can be symptomatic.
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_6
99
100
Although unusual, patients with Hodgkin lymphoma can present with superior vena cava syndrome, but chest pain, cough, and shortness of
breath are more common symptoms caused by a
large mediastinal mass. Lymphadenopathy found
only below the diaphragm is more common in
males and in elderly patients. Mesenteric lymphadenopathy is unusual in Hodgkin lymphoma.
Retroperitoneal lymphadenopathy can be painful,
but is more commonly asymptomatic and found on
a staging evaluation or as part of the investigation
to explain system symptoms such as fever, night
sweats, or weight loss. Epitrochlear lymph node
involvement is unusual in Hodgkin lymphoma.
Hodgkin lymphoma can involve essentially
any organ in the body as either a site of presentation or by spread from lymphatic involvement.
However, extranodal presentation of Hodgkin
lymphoma is unusual. The most common sites to
be involved are the spleen, liver, lungs, pleura,
and bone marrow, although Hodgkin lymphoma
confined to these sites is rare. Hodgkin lymphoma
can rarely present in unusual extranodal sites.
Primary CNS [1] and cutaneous [2] Hodgkin lymphoma are rare but well described. Perianal presentations are seen more commonly in patients
with HIV infection. Gastrointestinal system,
bone, genitourinary system, and other unusual
sites are extremely rare but have been described.
Bone involvement can be seen as an “ivory vertebrae,” i.e., a densely sclerotic vertebrae [3].
By far the most common systemic symptoms
that occur as the presenting manifestations of
Hodgkin lymphoma are fevers, night sweats,
weight loss, pruritus, and fatigue. These occur in
a minority of patients but can present diagnostic
challenges. Hodgkin lymphoma is one of the illnesses that can cause fever of unknown origin.
Occasionally the fevers of Hodgkin lymphoma
occur intermittently with several days of fevers
alternating with afebrile periods. This is the Pel-­
Ebstein fever [4, 5] that is rare, but typically
occurs in the evening. Fevers from Hodgkin lymphoma can be prevented with nonsteroidal anti-­
inflammatory drugs such as naproxen [6].
The presence of drenching night sweats (i.e.,
as opposed to dampness of the head and neck)
J. O. Armitage and J. W. Friedberg
and unexplained weight loss are both characteristics of Hodgkin lymphoma and, along with
fever, are associated with a poor prognosis.
Pruritus can be the presenting manifestation of
Hodgkin lymphoma. Such patients sometimes
have severely excoriated skin and sometimes
have been diagnosed as having neurodermatitis.
Patients who present with refractory pruritus are
often grateful to find the explanation of their
symptoms which usually disappear with the initiation of therapy. As with other lymphomas,
fatigue can be an important, although nonspecific, symptom and also usually improves with
therapy. There are many unusual, but welldescribed, presentations for Hodgkin lymphoma. One rare but very characteristic
presentation is alcohol-induced pain [7, 8]. The
pain typically begins soon after drinking alcohol
and occurs primarily in areas of involvement by
lymphoma. The pain can be quite severe and last
for variable periods of time. Patients with the
symptom have often discontinued alcohol
before the diagnosis of Hodgkin lymphoma, and
to elicit the symptom often requires specific
questioning by the physician.
Patients can present with Hodgkin lymphoma
involving the skin, but cutaneous abnormalities
are more often paraneoplastic phenomenon.
These can include erythema nodosum [9]; ichthyosiform atrophy [10]; acrokeratosis paraneoplastica [11]; granulomatous slack skin [12];
nonspecific urticarial, vesicular, and bullous
lesions [13]; and others.
A variety of other unusual presentations of
Hodgkin lymphoma have been reported. Patients
can present with nephrotic syndrome [14], symptoms of hypercalcemia [15–17], jaundice due to
cholestasis without involvement of the liver by
the lymphoma, and the “vanishing bile duct syndrome” [18, 19].
Hodgkin lymphoma very rarely presents with
a primary tumor in the CNS causing the symptoms of a brain tumor characteristic of the site of
involvement. Other neurological manifestations
that can be present at the diagnosis of Hodgkin
lymphoma involve a variety of paraneoplastic
syndromes. These include paraneoplastic
6
Clinical Evaluation
c­erebellar degeneration [20], which typically
presents with ataxia, dysarthria, nystagmus, and
diplopia. The symptoms may precede the diagnosis of Hodgkin lymphoma by many months.
Hodgkin lymphoma can, of course, present with
spinal cord compression from retroperitoneal
and osseous tumors. Other rare manifestations
include limbic encephalitis (i.e., which presents
with memory loss and amnesia), peripheral neuropathy, and others.
6.2
Physical Findings
and Laboratory
Abnormalities
By far the most common physical findings in
Hodgkin lymphoma are enlarged lymph nodes
that might be in any lymph node-bearing area.
The lymph nodes are typically firm (i.e., “rubbery”) and vary from barely palpable to large
masses. However, almost any aspect of the physical examination can be made abnormal by the
presence of Hodgkin lymphoma. This might
include icterus, involvement of Waldeyer’s ring,
findings of superior vena cava syndrome, a sternal or suprasternal mass from tumor growing out
of the mediastinum, findings of a pleural effusion
or pericardial fusion, an intra-abdominal mass,
hepatomegaly or splenomegaly, skin involvement, and, rarely, cutaneous or neurological
abnormalities.
Almost any laboratory test can be abnormal at
the time of diagnosis of Hodgkin lymphoma, but
certain tests are characteristic and should be specifically evaluated. Patients can have leukocytosis or leukopenia. Neutrophilia and lymphopenia
are sometimes seen and can confer a poor prognosis. Eosinophilia can be found incidentally
before the diagnosis of Hodgkin lymphoma, and
Hodgkin lymphoma should always be included
in the differential diagnosis of unexplained eosinophilia [21]. In some cases, the explanation of the
eosinophilia is related to production of interleukin-­5 by the tumor cells [22, 23].
The most common hematological manifestation of Hodgkin lymphoma is anemia. The most
101
usual explanation seems to be a normocytic anemia associated with the presence of the tumor
that resolves after therapy. However, patients can
also have autoimmune hemolytic anemia [24]
and a microangiopathic hemolytic anemia as part
of the syndrome of thrombotic thrombocytopenic
purpura has been reported.
Patients can present with thrombocytopenia
for a variety of reasons including hypersplenism
and bone marrow involvement. However, idiopathic thrombocytopenic purpura can be a presenting manifestation of the disease [25].
Other rare hematological manifestations of
Hodgkin lymphoma have included autoimmune
neutropenia [26], hemophagocytic syndrome
[27], coagulation factor deficiencies [28], and
unexplained microcytosis [29].
Routine chemistry screening should be done in
patients with Hodgkin lymphoma and might
reveal renal or hepatic dysfunction, protein abnormalities, hypercalcemia, and hyperuricemia.
Elevated erythrocyte sedimentation rate and
C-reactive protein are frequently seen and have
been associated with a poor prognosis.
6.3
Pathologic Diagnosis:
The Biopsy
The oncologist must be certain that the Hodgkin
lymphoma diagnosis was based on an adequate
biopsy specimen that was examined using appropriate morphologic and immunohistochemical
criteria. Whole lymph node excision is preferable for pathologic examination. The pathologic
diagnosis of Hodgkin lymphoma is fully discussed in Chap. 3.
The site of biopsy must be determined with
the radiologist and surgeon. In general, the largest abnormal peripheral lymph node should be
excised. If a fluorine-18-deoxyglucose positron
emission tomography (FDG-PET) has been performed, the patient should be biopsied in the
most avid site to avoid a partially necrotic zone.
If there are only deep nodes, the following
types of biopsy can be proposed. A thoracoscopic or laparoscopic approach under general
J. O. Armitage and J. W. Friedberg
102
anesthesia with, if necessary, preoperative
localization to facilitate resection can be performed [30]. Image-­guided core needle biopsy
is increasingly used and has a rising success
rate of more than 90% [31–33]. However, the
method has the disadvantage of only permitting
relatively small biopsies. In addition, this type
of biopsy is capable of sampling several core
specimens with a single biopsy tract. Largevolume cutting needles, ranging from 18 to
14 G, yield enough tissue for most immunochemistry stainings and even for RNA extraction from frozen tissue (Fig. 6.1). Fine-­needle
aspiration cytology should not be used for diagnosis of Hodgkin lymphoma, but may help in a
screening procedure, before biopsy [34].
Several pathologic pitfalls or differential
diagnoses should be kept in mind. Drugs such
as phenytoin or antibiotics may cause histologic changes within lymph nodes that may
mimic Hodgkin lymphoma, particularly the
mixed cellularity subtype. Other benign conditions like infectious mononucleosis, lymphoid
CD15
hyperplasia, or Castleman disease may produce
lymphadenopathy with histologic features similar to those of Hodgkin lymphoma. In fact, the
distinction between different diseases, including certain forms of non-Hodgkin lymphoma
(NHL), has been made clearer, thanks to a better definition of the entities by the WHO classification. T-cell-rich large B-cell lymphoma is
usually included in the differential diagnoses of
both nodular lymphocyte-­predominant Hodgkin
lymphoma and classical Hodgkin lymphoma,
while anaplastic CD30-positive NHL may display similar histology to that of classical
Hodgkin lymphoma. Nevertheless, molecular
studies require adequate material, sometimes
including frozen tissue in difficult cases, and
the role of the clinician is to make sure that the
node to be analyzed is given to an experienced
laboratory. If the clinical presentation of disease is not typical for the given pathologic
diagnosis, then a review of the pathology by an
expert hematopathologist should be considered
or even a second biopsy.
CD30
Fig. 6.1 Core needle biopsy for Hodgkin lymphoma with immunostainings for CD15 and CD30
6
Clinical Evaluation
6.4
103
taging Systems for Hodgkin
S
Lymphoma
The initial clinical evaluation and staging of
patients with Hodgkin lymphoma serve to confirm the Hodgkin lymphoma diagnosis, determine the extent and distribution of disease,
evaluate the patient’s fitness for standard treatments, and provide prognostic information
(Table 6.1).
Several staging systems were developed very
early and modified according to the progress
made in imaging and treatment of the disease.
The Ann Arbor staging was developed in the
Table 6.1 Lugano classification
Revised staging system for primary nodal lymphomas
Stage
Involvement
I
II
One node or a group of adjacent nodes
Extranodal (E)
status
Single
extranodal
lesions
Stage I or II
Two or more nodal groups on the same side of
the diaphragm
II bulkya
II as above with “bulky” disease
Not acceptable
Advanced III
Nodes on both sides of the diaphragm, nodes
Not acceptable
above the diaphragm with spleen involvement
IV
Additional noncontiguous extralymphatic
Not acceptable
involvement
NOTE: Extent of disease is determined by positron emission tomography for avid lymphomas and computed
tomography for non-avid histologies. Tonsils, Waldeyer’s ring, and spleen are considered nodal tissue
a
Whether stage II bulky disease is treated as limited or advanced disease may be determined by histology and a
number of prognostic factors
Criteria for involvement of site
Tissue site
Clinical
FDG avidity
Test
Positive finding
Lymph nodes
Palpable
FDG-avid
PET-CT
Increased FDG
histologies
CT
PET-CT
Diffuse uptake
Spleen
Palpable
FDG-avid
CT
with SUV >
histologies
liver, solitary
Non-avid disease
mass, miliary
lesions, nodules
>13 cm
PET-CT
Diffuse uptake,
Liver
Palpable
FDG-avid
CT
mass nodules
histologies
Non-avid disease
Mass lesion(s),
CNS
Signs,
CT
leptomeningeal
symptoms
MRI
infiltration, mass
CSF
assessment lesions,
cytology, flow
cytometry
Other (leg, skin, lung, GI tract, bone, bone
Site
PET-CTb,
Lymphoma
marrow)
dependent
biopsy
involvement
CSF cerebrospinal fluid, CT computed tomography, FDG fluorodeoxyglucose, MRI magnetic resonance imaging,
PET positron emission tomography
b
PET-CT is adequate for determination of bone marrow involvement and can be considered highly suggestive for
involvement of other extralymphatic sites. Biopsy confirmation of those sites can be considered if necessary
(continued)
104
J. O. Armitage and J. W. Friedberg
Table 6.1 (continued)
Revised criteria for response assessment
Response and site
Complete
Lymph nodes and extralymphatic sites
Non-measured
Organ enlargement
New lesions
Bone marrow
Partial
Lymph nodes and extralymphatic sites
Non-measured lesions
Organ enlargement
New lesions
Bone marrow
No response or stable disease
Target nodes/nodal masses, extranodal lesions
Non-measured lesions
Organ enlargement
New lesions
Bone marrow
Progressive disease
Individual target nodes/nodal masses
Extranodal lesions
PET-CT-based response
Complete metabolic response
Score 1, 2, or 3c with or without a residual mass on
5PS+
It is recognized that in Waldeyer’s ring or extranodal
sites with high physiologic uptake or with activation
within the spleen or marrow (e.g., with chemotherapy
or myeloid colony-stimulating factors), uptake may
be greater than normal in the mediastinum and/or
liver. In this circumstance, complete metabolic
response may be inferred if uptake at sites of initial
involvement is no greater than the surrounding
normal tissue even if the tissue has high physiologic
uptake
Not applicable
Not applicable
None
No evidence of FDG-avid disease in marrow
Partial metabolic response
Score of 4 or 5+ with reduced uptake compared with
baseline and residual mass(es) of any size
At interim, these finding suggest responding disease
At the end of treatment, these finding indicate
residual disease
Not applicable
Not applicable
None
Residual uptake is higher than the uptake in normal
marrow but reduced compared with baseline (diffuse
uptake compatible with reactive changes from
chemotherapy allowed). If there are persistent focal
changes in the marrow in the context of a nodal
response, consideration should be given to further
evaluation with MRI or biopsy or an interval scan
No metabolic response
Score 405 with no significant change in FDG uptake
from baseline at interim or end of treatment
Not applicable
Not applicable
None
No change from baseline
Progressive metabolic disease
Score 4 or 5 with an increase in intensity or uptake
from baseline
New FDG-avid foci consistent with lymphoma at
interim or end of treatment assessment
6
Clinical Evaluation
105
Table 6.1 (continued)
Non-measured lesions
New lesions
None
New FDG-avid foci consistent with lymphoma rather
than another etiology (e.g., leg infection,
inflammation). If uncertain regarding etiology of new
lesions, biopsy or internal scan may be considered
Bone marrow
New or recurrent FDG-avid foci
5PS 5-point scale, CT computed tomography, FDG fluorodeoxyglucose, IHC immunohistochemistry, LDi longest
transverse diameter of a lesion, MRI magnetic resonance imaging, PET positron emission tomography, PPD cross
product of the LDi and perpendicular diameter, SDi shortest axis perpendicular to the LDi, SPD sum of the product
of the perpendicular diameters for multiple lesions
_A score of 3 in many patients indicates a good prognosis with standard treatment, especially if at the time of an
interim scan. However, in trials involving PET where de-escalation is investigated, it may be preferable to consider
a score of 3 as inadequate response (to avoid undertreatment). Measured dominant lesions: Up to six of the largest
dominant nodes, nodal masses, and extranodal lesions selected to be clearly measurable in two diameters. Nodes
should preferably be from disparate regions of the body and should include, where applicable, mediastinal and
retroperitoneal areas. Non-nodal lesions include those in solid organs (e.g., liver, spleen, kidneys, lungs), GI
involvement, cutaneous lesions, or those noted on palpation. Non-measured lesions: Any disease not selected as
measured, dominant disease and truly assessable disease should be considered not measured. These sites include
any nodes, nodal masses, and extranodal sites not selected as dominant or measurable or that do not meet the
requirements for measurability but are still considered abnormal, as well as truly assessable disease, which is any
site of suspected disease that would be difficult to follow quantitatively with measurement, including pleural
effusions, ascites, bone lesions, leptomeningeal disease, abdominal masses, and other lesions that cannot be
confirmed and followed by imaging. In Waldeyer’s ring or in extranodal sites (e.g., GI tract, liver, bone marrow),
FDG uptake may be greater than in the mediastinum with complete metabolic response, but should be no higher
than the surrounding normal physiologic uptake (e.g., with marrow activation as a result of chemotherapy or
myeloid growth factors)
c
PET 5PS: 1, no uptake above background; 2, uptake _ mediastinum; 3, uptake _ mediastinum but _ liver; 4, uptake
moderately _ liver; 5, uptake markedly higher than the liver and/or new lesions; X, new areas of uptake unlikely to
be related to lymphoma
JCO 2014 [37]
1970s, when radiotherapy was the main curative
treatment option, and was based on the tendency
of Hodgkin lymphoma to spread to contiguous
lymph nodes [35].
Since the Ann Arbor staging, several significant changes in the management of Hodgkin
lymphoma have taken place. The Cotswolds
modification of the Ann Arbor staging system
was introduced in 1989 to approve the use of
CT scanning for the detection of intra-abdominal disease, to formalize a definition of disease
bulk, and to provide guidelines for evaluating
the response to treatment [36]. The current
standard is the Lugano classification which
addresses both staging and re-staging
(Table 6.1) [37].
A prognostic factor score for advanced
Hodgkin lymphoma treated by chemotherapy has
been worked out, based mostly on biological
parameters, including serum albumin <4 g/dL,
hemoglobin <10.5 g/dL, male sex, stage IV disease, age >45 year, white cell count >15,000 mm−3,
and lymphocyte count <600 mm−3 [38].
These prognostic factors are used to define
risk-adapted therapy. However, as combined
modality treatment with modern chemotherapy
has become the standard procedure for patients
with early-stage disease, the risk of relapse is
reduced, and some of these factors are no longer
associated with a high risk of relapse. In addition,
computed tomography (CT) and fluorine-18-­
deoxyglucose positron emission tomography
(FDG-PET) are now routinely used for the staging and evaluation of the response to treatment.
PET-CT provides reliable information on treatment efficacy.
J. O. Armitage and J. W. Friedberg
106
6.5
Imaging Evaluation
of the Extent of Disease
and early-stage extranodal disease. Thus, even
experienced oncologists vary in their stage
assignment of patients with nearby but disconThanks to the progress and availability of imag- tinuous extranodal involvement [43]. However,
ing techniques, it has been possible to improve the involvement of two or more noncontiguous
the accuracy of clinical staging, so that invasive extranodal sites should typically be considered
pathologic procedures are rarely necessary. At indicative of stage IV disease. The use of risk-­
present, the established radiological technique adapted treatment with chemotherapy has
for the diagnosis of Hodgkin lymphoma is FDG-­ reduced the importance of such factors.
The definition of bulk has varied considerPET [39].
FDG-PET is based on the increased glycolysis ably in the literature. For the mediastinum, one
of cancer cells. This is visualized using the radio- definition involved measuring the greatest
active glucose analog FDG, which after phos- transverse diameter of the mediastinal mass on
phorylation is metabolically trapped within the a standard posteroanterior chest radiograph and
cell. Thus, FDG-PET has become an established dividing it by the maximal diameter of the chest
imaging modality to stage, restage, and monitor wall at its pleural surfaces, usually at the level
therapy and detect recurrent lymphoma. PET and of the diaphragm or alternatively at the T5–T6
CT, which, respectively, supply metabolic and interspace (Cotswolds approach) [36]. A ratio
anatomic information, are complementary, and exceeding one third (1:3) was considered bulky
interpretation of the PET portion of the study is and a negative feature among patients treated
more accurate when the results of PET correlate with RT alone or chemotherapy alone. There
with those of CT [40, 41]. Therefore, integrated are no widely accepted criteria for the definiPET-CT systems were developed which are now tion of bulk using measurements obtained from
the standard care [42]. If PET-CT is not available, CT scans: the Cotswolds Committee recoman alternative imaging technique is computed mended that to constitute bulk, a nodal mass
tomography (CT) scan of the neck, chest, abdo- must be greater than 10 cm in diameter [36],
men, and pelvis. In rare cases where it is desir- whereas in some ongoing trials, bulk was
able to avoid radiation exposure (such as defined as confluent nodal masses greater than
pregnancy), MRI may be utilized as a substitute 7 cm [44]. The Lugano criteria states that a single nodal mass, in contrast to multiple smaller
for CT imaging.
It is important that imaging results be inter- nodes, of 10 cm or greater than a third of the
preted within the framework of the known pat- transthoracic diameter at any level of thoracic
terns of spread and other prognostic factors. A vertebrae as determined by CT is the definition
certain degree of variation in the size of medias- of bulky disease for HL. A chest X-ray is not
tinal and hilar nodes is normal, but those measur- required to determine bulk because of its high
ing more than 10 mm on the shortest cross section concordance with CT [37].
It appears that FDG-PET can largely eliminate
can be considered abnormal. However, although
the
necessity for doing bone marrow biopsies in
clearly abnormal findings on CT scanning may
patients
with Hodgkin lymphoma. One report of
be indicative of Hodgkin lymphoma, there is a
454
patients
found that no patients with a FDG-­
risk of false positives, particularly in the abdomen, when interpreting these findings. Therefore, PET scan assigned stage of 1 or 2 had a positive
when lymph nodes in the 15–20 mm range are bone marrow biopsy. The presence of focal skelseen, uptake on FDG-PET-CT is indicative of etal FDG-PET scan lesions identified positive
and negative bone marrow biopsies with a sensiinvolvement by lymphoma.
Substantial variations in stage assignment tivity and specificity of 85% and 86%. A negative
have nevertheless been demonstrated among FDG-PET scan for skeletal lesions had a 99%
patients with extranodal involvement, specifi- negative predictive value for the results of a bone
cally regarding the distinction between stage IV marrow biopsy [37, 45].
6
Clinical Evaluation
6.6
linical Evaluation During
C
Therapy
Clinical evaluation during treatment can be an
important component of the individualization of
treatment intensity. A rapid early response to initial therapy is increasingly recognized as a favorable prognostic factor among Hodgkin lymphoma
patients. Response can be evaluated by FDG-­
PET-­CT after two or three cycles of chemotherapy. Performing PET early during treatment has
also proved to be prognostically important and
has been incorporated into the response criteria.
Thus, a meta-analysis demonstrated that for lowto intermediate-risk Hodgkin lymphoma patients,
PET may be a good prognostic indicator after a
few cycles of standard chemotherapy [46].
FDG-PET scans performed after two cycles of
therapy are increasingly being utilized to guide
subsequent treatment [47–50]. The results of
interim FDG-PET scans have been used to
shorten the duration of therapy, to complete a
“standard” course of chemotherapy, to escalate
treatment to more intensive chemotherapy in
patients with a slow response, and to deescalate
therapy in patients with an excellent response.
6.7
Definition of the Response
to Treatment
FDG-PET scans have revolutionized determination of response to therapy in patients with
Hodgkin lymphoma. The often called “Lugano
criteria” have become the standard approach in
determining treatment response [51]. A key to
improving our ability to determine response to
therapy was standardization of interpretation of
PET scans. The so-called 5-point or Deauville
scoring system is recommended in the Lugano
criteria (Table 6.2) [37].
This system appears to have a high interobserver agreement. There has been debate about
what should be the definition of a complete
remission using the 5-point score. The consensus
appears to be a 5-point score of 3 or less is the
definition of a complete remission at the end of
therapy. Some studies of interim PET scans,
107
Table 6.2 5-point score (Deauville score)
Score
1
2
3
4
5
Definition
No uptake in sites of suspected lymphoma
Uptake but less than that seen in the
mediastinum
Uptake greater than seen in the mediastinum,
but less than seen in the liver
Uptake moderately higher than seen in the
liver
Uptake markedly higher than the liver and/or
new lesions
where the interim PET scan will be used to guide
possible treatment changes, have chosen to use a
more conservative 5-point score of 2 or less to
identify an early complete remission [47].
Ongoing studies are evaluating other criteria,
such as total metabolic tumor volume change,
and SUV change over time which may have less
variability between observers.
6.8
Complete Remission
The patient has no clinical, radiologic, or other
evidence of Hodgkin lymphoma. Changes due to
the effects of previous therapy (i.e., radiation
fibrosis) may, however, be present.
The category (CRu) has been eliminated from
the updated response criteria and now denotes
patients whose remission status is unclear,
because they display no clinical evidence of
Hodgkin lymphoma, but some radiologic abnormality that persists at a site of previous disease. In
this respect, it is generally recognized that imaging abnormalities may persist following treatment
and do not necessarily signify active disease [52].
It must be borne in mind that after mediastinal RT, thymic rebound, reactive lymph node
hyperplasia, or subclinical radiation pneumonitis may lead to abnormalities on FDG-PET [53].
To avoid false-positive interpretations, some
authors recommend that FDG-PET re-evaluation
should be delayed until 3 months after the completion of mediastinal RT, although the characteristic appearance of post-RT lung changes
occurring before 3 months can usually be distinguished from lymphoma by experienced nuclear
radiographers [42].
J. O. Armitage and J. W. Friedberg
108
The inclusion of PET in the new response criteria and the removal of CRu have simplified the
management of lymphoma patients by removing
some of the limiting factors of CT, which include
the size of lymph nodes that indicates involvement, the differentiation of unopacified bowel
from lesions in the abdomen and pelvis, the
inability to distinguish viable tumor from
necrotic/fibrotic lesions after therapy, and the
characterization of small lesions. A combined
PET-CT scan with a Deauville score of 1, 2, or 3
is consistent with complete remission.
6.9
Follow-Up Management
The manner in which patients are evaluated after
completing treatment may vary according to
whether treatment was administered in a clinical
trial or clinical practice and whether it was delivered with curative of palliative intent. In a clinical
trial, the requirement of uniform reassessment
may lead to follow-up studies that would not be
routinely done in practice.
Follow-up should involve identifying relapse
but also focus on identifying and dealing with
long-term adverse effects of treatment. These can
include secondary cancers, cardiac toxicity, thyroid disease, depression, and fertility issues [54].
Good clinical judgment, careful recording of
history, and a thorough physical examination are
the most important components of monitoring
patients after treatment. A complete blood count,
selected serum chemistry studies, and a sedimentation rate are frequently done with each visit.
However, there is no evidence to support the need
for regular surveillance CT scans. The patient or
physician identifies the relapse in more than 80%
of cases without imaging studies [55]. The most
important potential reason to do surveillance
imaging would be the detection of early relapse
that allowed early institution of salvage therapy
and increased survival. However, there is no evidence to support this hypothesis. One study of
241 patients that compared patients treated at different centers who did or did not do routine surveillance imaging found a 97% overall survival
rate in patients who received routine surveillance
imaging and a 96% 5-year survival rate in patients
who were only followed clinically [56]. In both
groups, salvage therapy was effective with only
one patient in the routine surveillance imaging
group dying of Hodgkin lymphoma. It was calculated that each relapse detected by surveillance
imaging costs $629,615, with no benefit in eventual outcome. Similar results have been found in
the use of surveillance imaging in pediatric
Hodgkin lymphoma [57].
In addition to financial costs, surveillance
imaging has other “side effects.” One study found
that patients undergoing surveillance imaging
had increased anxiety and fear associated with
the images [58]. In addition, it is known that CT
scans deliver a high level of radiation and are a
significant cause of cancer [59, 60].
An alternative to using CT scans would be
the use of FDG-PET scans as a potential tool for
the detection of relapse. However, in a prospective study of 36 Hodgkin lymphoma patients,
routine FDG-PET correctly identified all 5
relapses that followed treatment, but had a falsepositive rate of 55% [61]. A more recent study
using PET-CT scans showed a positive predictive value of only 28% for routine PET-CT scans
for surveillance for relapse [62]. Given the
observation that patients with cHL who are
event-free at 2 years have an excellent outcome
regardless of baseline prognostic factors, surveillance imaging beyond 2 years has not been
demonstrated to have value [63].
6.10
Conclusion
The careful and accurate clinical evaluation of
patients with Hodgkin lymphoma from presentation to follow-up in remission has a significant
impact on treatment outcome. The ability to perform an excellent history and physical and
knowledge regarding when, where, and how to
perform laboratory evaluations, images, and
biopsies are necessary for excellent care.
References
1. Gerstner ER, Abrey LE, Schiff D, Ferreri AJ, Lister
A, Montoto S et al (2008) CNS Hodgkin lymphoma.
Blood 112(5):1658–1661
6
Clinical Evaluation
2. Tassies D, Sierra J, Montserrat E, Marti R, Estrach T,
Rozman C (1992) Specific cutaneous involvement in
Hodgkin’s disease. Hematol Oncol 10(2):75–79
3. Granger W, Whitaker R (1967) Hodgkin’s disease in
bone, with special reference to periosteal reaction. Br
J Radiol 40(480):939–948
4. Pel PK (1887) Pseudoleukämie oder chronisches
Rückfallsfieber?
Symptomatol
Sogenannten
Pseudoleukämie II:644
5. Ebstein W (1887) Das chronische Rückfallsfieber,
eine neue Infectionskrankheit. Berl Klin Wochenschr
24:565
6. Chang JC, Gross HM (1985) Neoplastic fever
responds to the treatment of an adequate dose of
naproxen. J Clin Oncol 3(4):552–558
7. Bichel J (1959) The alcohol-intolerance syndrome in
Hodgkin’s disease. Acta Med Scand 164(2):105–112
8. James AH (1960) Hodgkin’s disease with and without
alcohol-induced pain. A clinical and histological comparison. Q J Med 29:47–66
9. Simon S, Azevedo SJ, Byrnes JJ (1985) Erythema
nodosum heralding recurrent Hodgkin’s disease.
Cancer 56(6):1470–1472
10. Ronchese F, Gates DC (1956) Ichthyosiform atrophy of the skin in Hodgkin’s disease. N Engl J Med
255(6):287–289
11. Lucker GP, Steijlen PM (1995) Acrokeratosis paraneoplastica (Bazex syndrome) occurring with acquired
ichthyosis in Hodgkin’s disease. Br J Dermatol
133(2):322–325
12. Noto G, Pravata G, Miceli S, Arico M (1994)
Granulomatous slack skin: report of a case associated
with Hodgkin’s disease and a review of the literature.
Br J Dermatol 131(2):275–279
13. Milionis HJ, Elisaf MS (1998) Psoriasiform lesions
as paraneoplastic manifestation in Hodgkin’s disease.
Ann Oncol 9(4):449–452
14. Dabbs DJ, Striker LM, Mignon F, Striker G (1986)
Glomerular lesions in lymphomas and leukemias. Am
J Med 80(1):63–70
15. Rieke JW, Donaldson SS, Horning SJ (1989)
Hypercalcemia and vitamin D metabolism in
Hodgkin’s disease. Is there an underlying immunoregulatory relationship? Cancer 63(9):1700–1707
16. Seymour JF, Gagel RF (1993) Calcitriol: the major
humoral mediator of hypercalcemia in Hodgkin’s
disease and non-Hodgkin’s lymphomas. Blood
82(5):1383–1394
17. Laforga JB, Vierna J, Aranda FI (1994)
Hypercalcaemia in Hodgkin’s disease related to prostaglandin synthesis. J Clin Pathol 47(6):567–568
18. Lieberman DA (1986) Intrahepatic cholestasis due
to Hodgkin’s disease. An elusive diagnosis. J Clin
Gastroenterol 8(3 Pt 1):304–307
19. Hubscher SG, Lumley MA, Elias E (1993) Vanishing
bile duct syndrome: a possible mechanism for intrahepatic cholestasis in Hodgkin’s lymphoma. Hepatology
17(1):70–77
20. Hammack J, Kotanides H, Rosenblum MK, Posner
JB (1992) Paraneoplastic cerebellar degeneration.
109
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
II. Clinical and immunologic findings in 21 patients
with Hodgkin’s disease. Neurology 42(10):1938–1943
Reid TJ III, Mullaney M, Burrell LM, Redmond J III,
Mangan KF (1994) Pure red cell aplasia after chemotherapy for Hodgkin’s lymphoma: in vitro evidence
for T cell mediated suppression of erythropoiesis and
response to sequential cyclosporin and erythropoietin.
Am J Hematol 46(1):48–53
Samoszuk M, Nansen L (1990) Detection of interleukin-­
5 messenger RNA in reed-Sternberg cells
of Hodgkin’s disease with eosinophilia. Blood
75(1):13–16
Di Biagio E, Sanchez-Borges M, Desenne JJ,
Suarez-Chacon R, Somoza R, Acquatella G (1996)
Eosinophilia in Hodgkin’s disease: a role for interleukin 5. Int Arch Allergy Immunol 110(3):244–251
Bjorkholm M, Holm G, Merk K (1982) Cyclic
autoimmune hemolytic anemia as a presenting
manifestation of splenic Hodgkin’s disease. Cancer
49(8):1702–1704
Kirshner JJ, Zamkoff KW, Gottlieb AJ (1980)
Idiopathic thrombocytopenic purpura and Hodgkin’s
disease: report of two cases and a review of the literature. Am J Med Sci 280(1):21–28
Heyman MR, Walsh TJ (1987) Autoimmune neutropenia and Hodgkin’s disease. Cancer 59(11):1903–1905
Kojima H, Takei N, Mukai Y, Hasegawa Y, Suzukawa
K, Nagata M et al (2003) Hemophagocytic syndrome
as the primary clinical symptom of Hodgkin’s disease. Ann Hematol 82(1):53–56
Slease RB, Schumacher HR (1977) Deficiency of coagulation factors VII and XII in a patient with Hodgkin’s
disease. Arch Intern Med 137(11):1633–1635
Shoho AR, Go RS, Tefferi A (2000) 22-Year-old
woman with severe microcytic anemia. Mayo Clin
Proc 75(8):861–864
De Kerviler E, Gossot D, Frija J (1996) Localization
techniques for the thoracoscopic resection of pulmonary nodules. Int Surg 81(3):241–244
de Kerviler E, Guermazi A, Zagdanski AM, Meignin
V, Gossot D, Oksenhendler E et al (2000) Image-­
guided core-needle biopsy in patients with suspected
or recurrent lymphomas. Cancer 89(3):647–652
Picardi M, Gennarelli N, Ciancia R, De Renzo A,
Gargiulo G, Ciancia G et al (2004) Randomized
comparison of power Doppler ultrasound-directed
excisional biopsy with standard excisional biopsy
for the characterization of lymphadenopathies in
patients with suspected lymphoma. J Clin Oncol
22(18):3733–3740
Agid R, Sklair-Levy M, Bloom AI, Lieberman S,
Polliack A, Ben-Yehuda D et al (2003) CT-guided
biopsy with cutting-edge needle for the diagnosis of
malignant lymphoma: experience of 267 biopsies.
Clin Radiol 58(2):143–147
Landgren O, Porwit MacDonald A, Tani E, Czader
M, Grimfors G, Skoog L et al (2004) A prospective
comparison of fine-needle aspiration cytology and
histopathology in the diagnosis and classification of
lymphomas. Hematol J 5(1):69–76
110
35. Rosenberg SA, Boiron M, DeVita VT Jr, Johnson
RE, Lee BJ, Ultmann JE et al (1971) Report of the
committee on Hodgkin’s disease staging procedures.
Cancer Res 31(11):1862–1863
36. Lister TA, Crowther D, Sutcliffe SB, Glatstein E,
Canellos GP, Young RC et al (1989) Report of a committee convened to discuss the evaluation and staging
of patients with Hodgkin’s disease: Cotswolds meeting. J Clin Oncol 7(11):1630–1636
37. Cheson BD, Fisher RI, Barrington SF, Cavaoli
F, Schwarz LH, Zucca E, Lister TA (2014)
Recommendations for initial evaluation, staging, and
response assessment of Hodgkin and non-Hodgkin
lymphoma: the Lugano classification. J Clin Oncol
32(27):3059–3067
38. Hasenclever D, Diehl V (1998) A prognostic score for
advanced Hodgkin’s disease. International prognostic factors project on advanced Hodgkin’s disease. N
Engl J Med 339(21):1506–1514
39. Armitage JO (2005) Staging non-Hodgkin lymphoma.
CA Cancer J Clin 55(6):368–376
40. Blodgett TM, Meltzer CC, Townsend DW (2007) PET/
CT: form and function. Radiology 242(2):360–385
41. von Schulthess GK, Steinert HC, Hany TF (2006)
Integrated PET/CT: current applications and future
directions. Radiology 238(2):405–422
42. Kazama T, Faria SC, Varavithya V, Phongkitkarun S,
Ito H, Macapinlac HA (2005) FDG PET in the evaluation of treatment for lymphoma: clinical usefulness
and pitfalls. Radiographics 25(1):191–207
43. Connors JM, Klimo P (1984) Is it an E lesion or stage
IV? An unsettled issue in Hodgkin’s disease staging. J
Clin Oncol 2(12):1421–1423
44. Laskar S, Gupta T, Vimal S, Muckaden MA, Saikia
TK, Pai SK et al (2004) Consolidation radiation after
complete remission in Hodgkin’s disease following
six cycles of doxorubicin, bleomycin, vinblastine,
and dacarbazine chemotherapy: is there a need? J Clin
Oncol 22(1):62–68
45. El-Galaly TC, d’Amore F, Mylam KJ et al (2012)
Routine bone marrow biopsy has little or no therapeutic consequence for positron emission tomography/computed tomography-staged treatment-naïve
patients with Hodgkin lymphoma. J Clin Oncol
30(36):4508–4514
46. Terasawa T, Lau J, Bardet S, Couturier O,
Hotta T, Hutchings M et al (2009) Fluorine-18-­
fluorodeoxyglucose positron emission tomography
for interim response assessment of advanced-stage
Hodgkin’s lymphoma and diffuse large B-cell
lymphoma: a systematic review. J Clin Oncol
27(11):1906–1914
47. Radford J, Illidge T, Counsell N, Hancock B,
Pettengell R, Johnson P, Wimperis J, Culligan D,
Popova B, Smith P, McMillan A, Brownell A, Kruger
A, Lister A, Hoskin P, O’Doherty M, Barrington
S (2015) Results of a trial of PET-directed therapy
for early-stage Hodgkin’s lymphoma. N Engl J M
372(17):1598–1607
J. O. Armitage and J. W. Friedberg
48. Raemaekers JMM, Andre MPE, Federico M, Girinsky
T, Oumedaly R, Brusamolino E, Brice P, Ferme C,
van der Maazen R, Gotti M, Bouabdallah R, Sebban
CJ, Lievens Y, Re A, Stamatoullas A, Morschhauser
F, Lugtenburg PJ, Abruzzese E, Olivier P, Casasnovas
RO, van Imhoff G, Raveloarivahy T, Bellei M, vad der
Borght T, Bardet S, Versari A, Hutchings M, Meignan
M, Fortpied C (2014) Omitting radiotherapy in early
positron emission tomography-negative stage I/II
Hodgkin lymphoma is associated with an increased
risk of early relapse: clinical results of the preplanned
interim analysis of the randomized DORTC/LYSA/
FIL H10 trial. J Clin Oncol 32(12):1188–1194
49. Press OW, Li H, Schoder H, Straus DJ, Moskowitz
CH, LeBlanc M, Rimsza LM, Bartlett NL, Evens
AM, Mittra ES, LaCasce AN, Sweetenham JW, Barr
PM, Fanale MA, Knopp MC, Noy A, Hsi ED, Cook
JR, Lechowicz MJ, Gascoyne RD, Leonard JP, Kahl
BS, Cheson BD, Fisher RI, Friedbert JW (2016) US
intergroup trial of response-adapted therapy for stage
III to IV Hodgkin lymphoma using early interim
fluorodeoxyglucose-­
positron emission tomography
imaging: southwest oncology group S0816. J Clin
Oncol 34(17):2020–2027
50. Johnson P, Federico M, Kirkwood A, Fossa A,
Berkahn L, Carella A, d’Amore F, Englad G,
Franceschetto A, Fulham M, Luminari S, O’Doherty
M, Patrick P, Roberts T, Sidra G, Stevens L, Smith
P, Trotman J, Viney Z, Radford J, Barrington S
(2016) Adapted treatment guided by interim PET-CT
in advanced Hodgkin’s lymphoma. N Engl J Med
374(25):2419–2429
51. Barrington SF, Mikhaeel NG, Kostakoglu L, Meignan
M, Hutchings M, Mueller SP, Schwartz LH, Zucca
EM, Fisher RI, Trotman J, Hoekstra OS, Hicks RJ,
O’Doherty MJ, Hustinx R, Biggi A, Cheson BD (2014)
Role of imaging in the staging and response assessment of lymphoma: consensus of the International
Conference on Malignant Lymphomas Imaging
Working Group. J Clin Oncol 32(27):3048–3058
52. Radford JA, Cowan RA, Flanagan M, Dunn G,
Crowther D, Johnson RJ et al (1988) The significance
of residual mediastinal abnormality on the chest
radiograph following treatment for Hodgkin’s disease. J Clin Oncol 6(6):940–946
53. Jerusalem G, Hustinx R, Beguin Y, Fillet G (2005)
Positron emission tomography imaging for lymphoma. Curr Opin Oncol 17(5):441–445
54. Ng AK (2014) Current survivorship recommendations
for patients with Hodgkin lymphoma: focus on late
effects. Blood 124(23):3373–3379
55. Henry-Amar M, Friedman S, Hayat M, Somers R,
Meerwaldt JH, Carde P et al (1991) The EORTC
lymphoma cooperative group. Erythrocyte sedimentation rate predicts early relapse and survival
in early-stage Hodgkin disease. Ann Intern Med
114(5):361–365
56. Pingali SR, Jewell SW, Havlat L et al (2014) Limited
utility of routine surveillance imaging for classical
6
57.
58.
59.
60.
Clinical Evaluation
Hodgkin lymphoma patients in first complete remission. Cancer 120:2122
Voss SD, Chen L, Constine LS et al (2012)
Surveillance computed tomography imaging and
detection of replace in intermediate – and advancedstage pediatric Hodgkin’s lymphoma: a report
from the Children’s oncology group. J Clin Oncol
30(21):2635–2640
Thompson CA, Charlson ME, Schenkein E et al
(2010) Surveillance CT scans are a source of anxiety
and fear of recurrence in long-term lymphoma survivors. Ann Oncol 21(11):2262–2266
Berrington de Gonzalez A, Mahesh M, Kim KP et al
(2009) Projected cancer risks from computed tomographic scans performed in the United States in 2007.
Arch Intern Med 169(22):2071–2077
Smith-Bindman R, Lipson J, Marcus R et al (2009)
Radiation dose associated with common computed
tomography examinations and the associated life-
111
time attributable risk of cancer. Arch Intern Med
169(22):2078–2086
61. Jerusalem G, Beguin Y, Fassotte MF et al (2003) Early
detection of relapse by whole-body positron emission tomography in the follow-up of patients with
Hodgkin’s disease. Ann Oncol 14(1):123–130
62. El-Galay TC, Mylam KJ, Brown P et al (2012)
Positron emission tomography/computed tomography
surveillance in patients with Hodgkin lymphoma in
first remission has a low positive predictive value and
high costs. Haematologica 97(6):931–936
63. Hapgood G, Zheng Y, Sehn LH, Villa D, Klasa R,
Gerrie AS, Shenkier T, Scott DW, Gascoyne RD,
Slack GW, Parsons C, Morris J, Pickles T, Connors
JM, Savage KJ (2016) Evaluation of the risk of
relapse in classical Hodgkin lymphoma at event-free
survival time points and survival comparison with the
general population in British Columbia. J Clin Oncol
34(21):2493–2500
7
Functional Imaging in Hodgkin
Lymphoma
Andrea Gallamini, Bruce Cheson,
and Martin Hutchings
Contents
7.1
Introduction 114
7.2
History of Imaging in Hodgkin Lymphoma 115
7.3
7.3.1
7.3.2
Background of PET and the FDG Tracer asic Principles of PET B
The FDG Tracerc
115
115
116
7.4
7.4.1
7.4.2
7.4.3
7.4.4
7.4.4.1
7.4.4.2
7.4.5
7.4.6
7.4.7
7.4.8
PET in Clinical Management of Hodgkin Lymphoma taging S
Early Assessment of Chemosensitivity Final Response Assessment
Interpretation Criteria Interim PET Scan End-of-Treatment PET Scan Treatment Response Assessment to Immune Checkpoint Inhibitors PET in Radiotherapy Planning PET for Response Prediction During Salvage Treatment PET for Follow-Up of HL Patients in Complete Remission 116
116
118
119
121
121
121
122
122
125
126
7.5
7.5.1
7.5.2
7.5.3
7.5.4
PET Response-Adapted Therapy arly-Stage HL E
Advanced-Stage HL Post-chemotherapy PET/CT-Driven Consolidation Radiotherapy PET/CT-Adapted Therapy in Relapsed HL 127
127
128
129
129
7.6
7.6.1
7.6.1.1
7.6.1.2
7.6.1.3
7.6.2
7.6.3
Toward Revised Criteria for PET Scan Interpretation ualitative vs. Semiquantitative Assessment Q
From Anatomical to Functional Imaging Toward New Criteria for Response Assessment of New Drugs Biomarker Integration in Response Criteria Interim PET End-of-Treatment PET 130
130
130
131
131
131
132
A. Gallamini (*) · B. Cheson · M. Hutchings
Department of research and Clinical innovation,
Antoine Lacassagne Cancer Center, Nice, France
e-mail: andrea.GALLAMINI@nice.unicancer.fr
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_7
113
A. Gallamini et al.
114
7.7
FDG-PET/MRI
132
7.8
7.8.1
7.8.2
7.8.2.1
7.8.2.2
Future Perspectives
ther Tracers
O
Quantitative Methods for PET Reading (SUV, MTV, TLG)
Semiquantitative Assessment
Metabolic Tumor Volume
133
133
134
134
134
7.9
General Recommendations for the Use of PET in HL
135
References
7.1
Introduction
Hodgkin lymphoma (HL) has become a highly
curable malignancy with more than 90% of
patients alive and 80% considered cured after
long-term follow-up [1, 2]. Risk-adapted strategies have further led to improved outcomes for
high-risk patients, with less toxicity for low-risk
patients [3–7].
The staging of HL has undergone a major evolution. Imaging technology in this disease has
evolved over the past decades from the lymphangiogram to the intravenous pyelogram, ultrasound, liver-spleen radionuclide scan, computed
tomography (CT), and magnetic resonance imaging. Gallium scanning required a 2-day interval
between injection and scanning, which was not
clinically practical [8]. Today, positron emission
tomography combined with computed tomography (PET-CT) is the basis for staging and
response assessment [9–12]. Response assessment with CT uses changes in tumor size as the
main criterion. However, tumor shrinkage occurs
over time, and fibrosis in HL may take years following treatment to disappear. With these limitations, CT does not provide a sufficiently accurate
early assessment of response [13]. In contrast,
PET depends on tumor metabolism rather than
anatomy. Positron emission tomography using
[18F]-fluoro-2-deoxy-d-glucose
(FDG-PET)
provides an assessment of chemosensitivity when
performed early during standard treatment
[14–17].
In HL tissue, scattered neoplastic Hodgkin
Reed-Sternberg cells usually account for less
than 1% of the total cell count. Recently it has
been recognized that the malignant cells are able
135
to evade the immune system because of overproduction of the programmed death ligands (PDL)
PDL-1 and PDL-2 [18], despite the extensive
infiltration of lymph nodes by inflammatory
cells. The consequence is suppression of T-cell
activation and failure of immune recognition.
The relevance of this interaction is exemplified
by the efficacy of checkpoint inhibitors in HL
patients [19–21].
Functional imaging is critical for accurate
staging, restaging, and follow-up assessment of
patients with HL [9, 10]. Metabolic imaging is
currently performed using FDG-PET. This technology has become the most sensitive and specific technology, allowing us to better manage
HL patients [10]. A variety of clinical studies
have established the role of PET-CT scanning in
the risk-adapted management of HL patients,
leading to improved outcomes and reduced toxicity [3, 4, 22].
Despite major advances in patient outcome,
further refinements are warranted not only in the
interpretation of PET-CT but also in determining
how best to incorporate PET into patient management. In this chapter, we will review the history
of metabolic imaging in HL and its usefulness in
the staging and response assessment, including
its role in risk-adapted strategies allowing for
improved outcome for high-risk patients and
decreased toxicity for low-risk patients. We will
also discuss the current gaps in the application of
PET-CT, potential means to improve the current
response criteria, and speculate on the future of
metabolic imaging in the management of HL
patients. Newer and potentially more specific
PET tracers are also under investigation and will
be discussed in this chapter.
7
Functional Imaging in Hodgkin Lymphoma
7.2
History of Imaging
in Hodgkin Lymphoma
HL has been for long considered the paradigm for
tumor staging in oncology [23]. In the early
1970s, the Ann Arbor conference [23] and later
the Cotswold revised classification [24] introduced the concept that the disease spread per se is
able to identify distinct categories of patients with
different prognosis and treatment outcome.
Staging laparotomy using splenectomy and multiple nodal and organ biopsies including bone
marrow trephine biopsy (BMB) was originally
proposed as an accurate diagnostic means for
tumor staging [25]. This procedure had the merit
of shedding more light on the physiopathology of
tumor spread, becoming the “gold standard” to
assess sensitivity, specificity, and overall accuracy
of newly emerging radiological imaging [26].
In April 1970, during the Ann Arbor conference in Michigan, the concept of four-stage clinical staging (CS) was initially introduced to
distinguish patients staged with clinical and radiological means from those more accurately staged
with pathological staging (PS) [23]. Nonetheless,
the high accuracy of surgical staging was deemed
no longer necessary with the advent of active
multi-agent chemotherapy such as MOPP (mechlorethamine, vincristine, procarbazine, and prednisone). However, MOPP also led to sterilization
in more than 80% of patients treated [27]. Bipedal
lymphography and contrast-enhanced computed
tomography (ceCT) very soon superseded the
invasive procedures of PS including staging laparotomy, with the notable exception of BMB. The
latter became the only invasive technique used for
staging, as CT proved insufficient for evaluation
of HL infiltration in the bone marrow. Meanwhile,
the disease burden and the host reaction against
the tumor proved as the main prognostic parameters correlating with survival. This subsequently
provided the basis for a new classification of
prognostic factors in HL as (a) tumor related, (b)
host related, and (c) environment related [28],
providing the frame for three different risk groups
of HL patients with different long-term prognosis:
the early favorable, early unfavorable, and
advanced HL [29].
115
The clinical and radiological procedures of
HL staging, the definition of bulky nodal lesion,
and the nomenclature to define response to treatment have been described and standardized during the Cotswold meeting in 1989 [24]. In the
early nineties, another step forward in the tumor
staging accuracy was the use of functional imaging with 67Ga scintigraphy initially, and later with
positron emission tomography (PET), using the
glucose analogue 18F-fluorodeoxyglucose (FDG).
The latter was able to trace viable tumor tissue
selectively, thus resulting in the more accurate
diagnostic tool so far evaluable for lymphoma
staging and restaging. With the introduction of
integrated FDG-PET/CT scanners, unsuspected
nodal and extranodal lesions were detected,
which otherwise would have been missed by the
current diagnostic tools, including ceCT and
BMB [30]. FDG-PET/CT showed a better sensitivity and a similar specificity compared to FDG-­
PET stand alone and CeCT in detecting nodal
and extranodal disease [9, 10]. In HL and diffuse
large B-cell lymphoma (DLBCL), the bone and
bone marrow was by far the most frequently
detected extranodal site followed by the liver,
lung, and spleen. Due to the lower overall accuracy of BMB compared to FDG-PET/CT in
detecting bone or bone marrow, BMB was abandoned as the standard diagnostic tool to detect
BMI by HL and DLBCL [31, 32].
7.3
Background of PET
and the FDG Tracer
7.3.1
Basic Principles of PET
PET is a functional imaging modality based on
measurements of radiation from the decay of
positron-emitting radioactive nuclides. These
nuclides have excess protons which transform to
neutrons under the emission of positrons
(β+-decay). The positron randomly travels
2–3 mm in the tissue before it annihilates via collision with an electron, and thereby emitting two
photons (each 511 keV) at an angle very close to
180°. The two photons are registered by the ring
of scintillation detectors in the PET scanner. Two
A. Gallamini et al.
116
511 keV photons registered simultaneously (or
within a very narrow time frame) by two opposing detectors are considered a coincidence event
originating from positron annihilation. A PET
scanner holds several thousands of scintillation
detectors, organized in detector rings. The detector rings may be separated by leaded ring collimators (2D mode) in order to limit sources of
noise in the PET images. Data acquisition can be
either static or dynamic, and the data generated
provide both quantitative information and
images. The spatial resolution of PET is typically
around 3–5 mm, limited by the number of detectors and by the random travel of the positron [33].
The unstable positron-emitting isotopes used in
PET are produced by fusion of stable nuclei with
other particles. This is possible in a cyclotron, in
which the electrical repulsion between particles
is overcome by accelerating particles up to 30%
of the speed of light with a beam toward the target [34]. A radiochemistry laboratory is needed
to attach the isotopes to relevant tracer molecules.
The most common PET isotope molecules are
15
O, 13N, 11C, and 18F [35]. PET tracers of relevance to oncology target glucose metabolism,
hypoxia, blood flow, proliferation, amino acid
transport, protein synthesis, DNA synthesis,
apoptosis, and specific receptors.
Fusion PET/CT scanners incorporate the
hardware of high-resolution CT and PET into one
scanner, so that PET and CT as well as fusion
images are obtained in one scanning session.
PET/CT scanners have been available commercially since the late 1990s and have now replaced
the single-modality PET scanners. PET/CT provide anatomical localization of the FDG uptake,
as well as better distinction between pathological
findings and normal physiological uptake [36].
porter molecules (GLUT 1–5), which are overexpressed in cancer cells [38–40]. In the cell, FDG is
phosphorylated by hexokinase to FDG-6phosphate, which does not cross the cell membrane. Due to the low levels of
glucose-6-­
phosphatase in cancer cells and the
inability of FDG-6-phosphate to enter glycolysis,
the tracer is retained in the cancer cells [41].
Generally, the uptake of FDG is related to the
number of viable tumor cells [42, 43], but is dependent also on a number of physiological factors
including regional blood flow, blood glucose level,
and tissue oxygenation [44, 45]. FDG uptake is
very high in HL, but since the HRS cells only
make up a small fraction of the tumor volume, the
surrounding cells are accountable for most of the
increased FDG metabolism. FDG uptake is not
tumor specific and accumulates in a range of nonmalignant tissues, such as the brain, heart, and kidneys. Furthermore, activated inflammatory cells
take up FDG, which can cause false-positive
results in cancer imaging studies [46, 47]. This is
obviously important not only since HL patients
frequently experience infections but also because
chemotherapy and radiotherapy induce inflammatory responses in the tumor cells and the surrounding tissue. Increased tracer uptake is seen in
response to the early-phase tissue inflammation
induced by chemotherapy, while a “stunned” uptake
is observed immediately after therapy [48, 49].
7.4
ET in Clinical Management
P
of Hodgkin Lymphoma
7.4.1
Staging
Currently, the original Ann Arbor nomenclature
for HL staging is the standard tool to define disease prognosis and to guide treatment [23]. In
7.3.2 The FDG Tracer
the recent consensus statement of Lugano in
2014 for the use of FDG-PET for HL staging,
The
glucose
analogue
2-[18F]fluoro-2-­
the presence or absence of Systemic symproms
deoxyglucose (FDG) is the most versatile and the (A or B substage definition) was still considered
most widely used PET tracer. FDG is administered useful in HL to guide treatment as “B sympby intravenous injection. The use of FDG in tumor toms” were still considered adverse prognostic
imaging is based on Warburg’s finding that cancer indicators [9].
cells show accelerated glucose metabolism [37].
As previously mentioned, a committee of
FDG is transported into the cell via glucose trans- experts convened in 2014 so that FDG-PET/CT
7
Functional Imaging in Hodgkin Lymphoma
became the standard method to stage and restage
the vast majority of lymphomas, including HL [9,
10]. FDG-PET became a paradigmatic example
of the superiority of functional over classic radiological imaging for the following reasons: (1)
FDG-PET showed a higher sensitivity for nodal
staging; (2) FDG-PET was clearly superior in
detecting extranodal disease, both in the bone
marrow and in other organs; and (3) FDG-PET
had a consistent, large influence on HL staging,
with a potential impact on treatment strategy in a
substantial number of patients, and this became
even more evident upon introduction of integrated FDG-PET/CT scanners [50]. In an early
pioneer study using FDG-PET alone, CT and
fused FDG-PET/CT were prospectively compared in 99 newly diagnosed HL patients. In
nodal regions, the sensitivity of PET and PET/CT
was higher than that of CT (92% and 92% vs.
83%). FDG-PET had more false-positive nodal
sites than CT and FDG-PET/CT (1.6% vs. 0.7%
and 0.5%). For evaluation of organs, FDG-PET
and FDG-PET/CT had higher sensitivity (86%
and 73%), while CT detected only 37% of
involved organs. In conclusion, FDG-PET/CT
upstaged 19% and downstaged 5% of patients,
leading to a treatment modification in 9% of
patients [30]. A recent study prospectively compared a cohort of 96 HL patients [51]. Similar to
the previous study, radiologists and nuclear medicine physicians were blinded to the outcome of
the other modality and to the clinical course of
the patients. The breakdown of patients according to stage I to IV based on CT vs. FDG-PET/
CT was: 5 vs. 7, 49 vs. 37, 28 vs. 22, and 14 vs.
30, respectively. FDG-PET/CT changed the stage
in 33 (34%) patients, 28% were upstaged and 6%
downstaged. Upstaging was mainly caused by
detection of new extranodal involvement (47
sites in 26 patients) including the bone marrow,
spleen, and lung. Downstaging resulted from the
absence of FDG uptake in enlarged nodes
(<15 mm) in the abdomen and pelvis. FDG-PET/
CT led to a treatment modification in 20 (21%) of
the patients, with 16 patients being allocated to
more intensive treatment [51]. The role of bone
marrow infiltration (BMI) in stage upgrading was
even more evident in recent reports on the role of
FDG-PET/CT in staging of patients enrolled in
117
the prospective UK National Cancer Research
Institute RATHL clinical trial [52]. Out of 938
enrolled patients, FDG/PET-CT led to upstaging
in 159 patients (14%) and downstaging in 74
(6%). The most frequent staging migration was
from stage III to IV due to detection of disease
spread to extranodal sites (ENS), most frequently
in the bone marrow, lung, and others. In the cases
of discrepant results, follow-up images confirmed
the HL nature of the lesion detected by FDG-­
PET/CT only.
CT is insufficient for BMI evaluation in HL,
while PET/CT detects skeletal FDG uptake in
10–20% of patients [53, 54]. This observation
changed the perception that BMI is a rare occurrence in HL. Most studies using BMB for HL
detection in the bone report a BMI frequency of
only 5–8% [55]. The use of iliac crest bone marrow biopsy as a surrogate for the whole bone
marrow compartment has been challenged by
frequent finding of focal FDG lesions in the
bone marrow in patients undergoing PET/CT
staging. In addition, one-sided BMI has been
reported in nearly half of the HL patients undergoing bilateral bone marrow biopsies [56]. In a
recent large retrospective study performed by
the German Hodgkin Study Group, Voltin et al.
showed a PET/CT- detected BMI in 129/832
(15%) patients, while BMB was positive in only
20 (2%) [57]. With the gold-standard reference
of either a positive PET scan (which becomes
negative in subsequent follow-up scans) or a
positive BMB or both, the sensitivity, specificity, positive, and negative predictive values of
PET to detect BMI were 99.25%, 100%, 100%,
and 99.9%, respectively. In conclusion, PET/CT
has higher sensitivity for BMI than conventional
BMB [31, 57, 58]. Rare patients with an undetected BMI by PET/CT at baseline almost
exclusively present with advanced-stage disease
based on PET/CT. Thus, the added diagnostic
information from BMB very rarely leads to
changes in clinical management [31]. As far as
the pattern of FDG uptake is concerned, only
focal uptake, defined as a single spot of FDG
uptake at the bone level visible in at least two
PET/CT slices with an intensity of FDG
uptake ≥ liver, is considered as a harbinger of
BMI. On the other hand, patients with a diffuse
A. Gallamini et al.
118
FDG uptake (with an intensity ≥ liver) had disease outcomes identical to patients without any
FDG uptake [53, 54].
Finally, one of the most interesting technological progress in PET/CT is the measurement of
metabolic tumor volume (MTV) with more
sophisticated software counting all voxel contained in a single contoured tumor lesion. To overcome the so-called partial volume effect, only
voxel with an activity higher than a given threshold are counted. By multiplying the number of
counted voxels using a fixed coefficient, it is possible to calculate the MTV expressed in cubic centimeters. (Fig. 7.1) Upon identification of the best
cutoff value with ROC curve analysis, Cottereau et
al demonstrated that, in early stage HL, a MTV
value higher than 147 cm3 could single out a
smaller (46 vs. 157) patient subset with poor prognosis compared to unfavorable patients [59]: the
3-Y PFS was 71% vs. 84% [60]. Since the actual
standard of care for early-stage HL is tailored to
EORTC prognostic criteria or similar score systems, MTV could guide the treatment intensity of
early-stage HL. However, standardization problems due to intrinsic variability of PET-extracted
semiquantitative variables such as SUV (standardized uptake value), different protocols for patient
scanning and image acquisition/reconstruction, as
well as different thresholds of SUVmax still hamper the use of this biomarker for risk stratification
and treatment guidance in HL [61].
7.4.2
Early Assessment
of Chemosensitivity
Dimensional parameters providing readout of
tumor growth have extensively been used in standard radiological imaging to assess therapeutic
effects early during treatment using the so-called
RECIST criteria [62]. However, the kinetics of
tumor shrinkage and regrowth are not linear and
might overtake the prognostic impact of tumor
size variation. Patients with residual mass persisting at the end of treatment need longer follow-­up
for judging response to treatment [63–65]. FDGPET/CT is the ideal tool to assess viable neoplastic cells in the context of residual masses
detectable with CeCT. Since metabolic silencing
is immediately visible after chemotherapy [13],
response assessment with FDG-PET could be
performed during treatment, as early after one
[66, 67] or two [13, 15, 16, 68–70] cycles of chemotherapy, both in early- and advanced-stage HL.
Fig. 7.1 Metabolic tumor volume: calculation example and VOI drawing depending on software (By Kanoun et al.:
Plos One 2015;10(10):e0140830 (by permission))
7
Functional Imaging in Hodgkin Lymphoma
Despite the high FDG affinity of neoplastic
Reed-Sternberg cells attributed to the “Warburg
effect” of neoplastic tissue [37] or an m-TOR-­
mediated increase of transmembrane GLUT-1
protein [71], the high performance of interim
FDG-PET in predicting HL outcome probably
relies on the high affinity for the tracer of nonneoplastic microenvironment cells, which account for
more than 95% of the total cells present in HL
tissue [18]. Several publications stressed the role
of interim PET (iPET) in predicting ABVD treatment outcome in advanced-­stage HL [13, 15, 66–
70]. In a meta-analysis by Terasawa et al., interim
PET in advanced-stage HL showed a sensitivity
of 43–100% and specificity of 67–100%, respectively [70]. iPET proved to predict treatment outcome also in early stage, albeit with conflicting
results [16, 72–74] and a lower specificity and
positive predictive value [74].
Two questions concerning the ideal time for
early interim FDG-PET scanning are still unsettled: (1) what is the best time point for FDG-PET
after chemotherapy administration, and (2) what
is the ideal number of chemotherapy cycles before
the early interim FDG-PET scan? As far as the
point (1) is concerned, in mice undergoing FDG
scan, the FDG uptake by neoplastic cells and
reactive inflammatory macrophages was minimal
14 days after chemotherapy administration [49].
An earlier evaluation, immediately after chemotherapy, could coincide with the stunning of the
cellular glucose metabolism by immediate effects
of chemotherapy compromising the sensitivity of
the test [75]. In a review on interim FDG-PET
during early treatment, Kasamon concluded that
the optimal time for performing interim PET during chemotherapy ranges between 7 and 14 days
after chemotherapy [76]. The answer to point (2)
could depend on the aggressiveness of the tumor
and the efficacy of the chemotherapy. In HL, there
is most evidence for the use of FDG-PET after
two courses of chemotherapy. More recently, two
observational prospective studies have been published, stressing the good overall accuracy of
interim PET performed as early as after one single
cycle of chemotherapy, with very high sensitivity
and negative predictive value [66, 67] especially
in early-stage disease [67].
119
7.4.3
Final Response Assessment
Between 1999 and 2001, several papers reported
high sensitivity and specificity of FDG-PET in
tumor response assessment. In a meta-analysis of
13 studies on 408 HL patients, Zijlstra and colleagues demonstrated a pooled sensitivity and
specificity of PET in defining treatment outcomes
of 84% and 90%, respectively [77] (Fig. 7.2).
As a consequence, FDG-PET was proposed as
the essential tool for the definition of treatment
response in the majority of different FDG-avid
lymphoma, including HL. The most recent definitions of CR, PR, SD, or progressive disease were
integrated into the International Harmonization
Project for treatment response in lymphomas [78]
and the revised Lugano criteria [9]. The Deauville
5-point scale [79] has been used to quantify and
standardize the residual FDG uptake [10]. The
traditional concept of CR and CRu were abandoned, and the term complete metabolic response
(CMR) was proposed instead; similarly, the old
concept of partial remission (PR) based on dimensional criteria was replaced by partial metabolic
response (PMR) in FDG-­
avid lymphoma. No
metabolic response (NMR) was suggested for
nonresponse or stable disease patients, while progressive metabolic disease (PMD) was proposed
for clearly progressive patients. Traditionally,
dimensional criteria were maintained for the rare
non-FDG-avid lymphomas [9]. Preliminary
reports showed a better accuracy in defining treatment response and long-term outcome of the
5-point Deauville scale (5-PS) over the
International Harmonization Project (IHP) criteria, both in early and advanced HL [78, 80, 81].
Fallanca et al. in a small cohort of 101 patients
including 35 advanced HL cases were able to
show a better specificity, positive predictive value,
and overall accuracy of 5-PS over IHP criteria
(87% vs. 67%, 74% vs. 57%, and 86% vs. 76%,
respectively). Therefore, the reduced number of
false-positive outcomes in the end-of-treatment
PET proved to spare a significant number of
patients from unnecessary treatment.
Despite the good response to therapy, treatment
of HL can result in residual masses in up to 80% of
patients by conventional staging ­modalities, which
A. Gallamini et al.
120
Sensitivity (95% CI)
de Wit, 2001
Hueltenschmidt, 2001
Mikhaeel, 2000
Naumann, 2001
Spaepen, 2001
Weirauch, 2001
Zinzani, 1999
0
0.2
0.4
0.6
0.8
1
1,00 (0,69-1,00)
0,95 (0,74-1,00)
1,00 (0,29-1,00)
1,00 (0,03-1,00)
0,50 (0,19-0,81)
0,67 (0,30-0,93)
1,00 (0,29-1,00)
Pooled sensitivity = 0,84 (0,71 to 0,92)
Chi-square = 15,87; df = 6 (p = 0,0145)
Specificity (95% CI)
de Wit, 2001
Hueltenschmidt, 2001
Mikhaeel, 2000
Naumann, 2001
Spaepen, 2001
Weirauch, 2001
Zinzani, 1999
0,78 (0,56-0,93)
0,89 (0,72-0,98)
0,92 (0,62-1,00)
0,86 (0,71-0,95)
1,00 (0,93-1,00)
0,80 (0,56-0,94)
1,00 (0,69-1,00)
Pooled specificity = 0,90 (0,84 to 0,94)
Chi-square = 17,96; df = 6 (p = 0,0063)
0
0.2
0.4
0.6
0.8
1
Specificity
Fig. 7.2 Zijlstra unchanged. (a) Sensitivities and 95%
confidence intervals for studies assessing the diagnostic
accuracy of FDG-PET in patients with HD. (b) Specificity
and 95% confidence intervals for studies assessing the
diagnostic accuracy of FDG-PET in patients with HD.
(Asterisk) The diamond represents the 95% CI of the
pooled estimate (From Zijlstra 2006, by permission [77])
are most frequent in the anterior mediastinum [63,
65]. The original study by Jerusalem et al. [82]
showed the superior accuracy of FDG-­PET over
CeCT in predicting long-term disease outcome in
patients showing persisting FDG uptake. Many
reports confirmed the superior performance of
functional imaging in this setting. In 2008,
Terasawa et al. systematically reviewed all studies
published so far on this topic, reporting a pooled
sensitivity for HL ranging between 43% and
100%, with a specificity ranging between 67% and
100% [83]. In the pre-PET era, consolidation
radiotherapy had been used in advanced-­stage HL
showing radiological evidence of a residual
mass persisting after chemotherapy [2]. The endof-treatment PET was proposed to deliver consolidation radiotherapy (cRT) only in PET-positive
residual masses. As a consequence, the number of
patients undergoing cRT at the end of treatment
changed from 71% to 11% [22]. In the German
Hodgkin Study Group trial HD 15, cRT was
administered to patients having FDG-avid residual
mass of more than 2.5 cm at the end of chemotherapy. About 30% of the patients showed a residual mass, and in about 10% the residual mass
proved FDG avid. Comparing irradiated with nonirradiated patients, the 4-year PFS was 86.2% and
92.6%, respectively (p = 0.022). Overall, the NPV
at the end of therapy was 94% [22]. Savage et al.
reported very similar results in a retrospective
analysis of 163 advanced-stage HL patients undergoing cRT avter ABVD chemotherapy only in the
case of a residual FDG-avid mass of 2 cm or more
at the end of chemotherapy. Patients with a PETnegative scan (n = 130, 80%) had a 3-year time to
progression far superior to that of patients with a
7
Functional Imaging in Hodgkin Lymphoma
121
PET-positive scan (89% vs. 55%, p = 0.00001),
with no difference between those having bulky or
non-bulky disease. The NPV for end-of-treatment
PET was 92% [84]. However, NPV at the end-oftreatment PET depends on the efficacy of chemotherapy, being so low as 86% with low-intensity
regimen such as VEBEP [85].
7.4.4
Interpretation Criteria
7.4.4.1 Interim PET Scan
Early studies demonstrated that not all interim
PET scans could be dichotomized as positive or
negative; some patients had “equivocal” or
“inconclusive” results showing residual FDG
uptake defined as minimal residual uptake (MRU).
MRU definition evolved from residual unspecific
FDG uptake with intensity equal, or slightly superior to mediastinum [86], to an uptake with higher
intensity, equal to that of the liver [79]. This was
proposed with the aim to increase the specificity
and reduce false-positive results in predicting
treatment outcome [87]. These criteria have been
generally accepted by imaging specialists and clinicians as a useful tool for visual interim PET
assessment in lymphoma and have been incorporated in several guidelines and recommendation
by academic institutions [9, 10, 88–90] (Fig. 7.3).
The Deauville five-point scale (5-PS) has been
retrospectively and prospectively validated in HL
in four different studies [11, 12, 52, 81].
Prospective testing between national imaging
core laboratories also showed good interobserver
agreement in HL [91]. In standard treatment or
Fig. 7.3 The Deauville
5-point scale
(unchanged) (From
Meignan M, Leuk
Lymphoma 2009, by
permission [79];
Barrington S: Eur
J. Nucl Med Mol
Imaging
2010;37:1824–1833)
treatment intensification planned for interim
PET-positive patients, a Deauville score (DS) of
3 represents a complete metabolic response
(CMR). DS 3 also proved to be the most reproducible threshold when interpreting PET scans in
lymphoma patients [52]. If treatment is to be de-­
escalated, a more sensitive threshold is preferred
to avoid the risk of undertreatment [3, 7]. The use
of a higher threshold for a positive interim PET
scan was supported by the RATHL trial showing
that the DS in PET-negative patients (DS 1 vs.
DS 2 vs. DS 3) did not influence PFS, suggesting
that DS 3 is likely to represent CMR with standard ABVD treatment [4]. To obtain more stable
measures of residual activity in interim PET, the
Leipzig/Germany group proposed a semiquantitative method for interim PET reading, using the
SUVpeak of the residuum and the SUVmean of
the liver rather than SUVmax [92]. Since the
residual lesion detected in interim PET is often
very small and a VOI of 1 ml too large, the
SUVpeak was defined as the average value in the
maximum SUV voxel and the three hottest adjacent ones [92].
7.4.4.2 End-of-Treatment PET Scan
The presence of a FDG-avid residual mass at the
end of treatment represents a true diagnostic
dilemma. As mentioned above, only half of the
residual masses at the end of treatment in HL
have been considered as a harbinger of persisting
disease [93]. Primary mediastinal B-cell lymphoma is a rare disease often presenting in young
females with isolated mediastinal bulky lesion
and a good treatment outcome [94]. The disease
Deauville score (DS)
Score 1 no uptake
Score 2 uptake ≤ mediastinum
Score 3 uptake > mediastinum but ≤ liver
Score 4: moderately ↑ uptake > liver
Sensitive threshold
Specific threshold
Score 5 markedly ↑ uptake > liver and/or new sites of
disease
Barrington S: Eur J. Nucl Med Mol Imaging 2010;37:1824–1833
Meignan M. Leukemia & Lymphoma 2009;50(8):1257–1260
122
A. Gallamini et al.
is biologically and clinically related to nodular 2. Appearance of new lesions or growth of one
or more existing lesion(s) ≥ 50% at any time
sclerosis HL, where the putative normal cell
during treatment, occurring in the context of
counterpart is a thymic B cell [95]. End-of-­
lack of overall progression (<50% increase)
treatment PET scan in this disease is often associof overall tumor burden, as measured by SPD
ated with false-positive results [96, 97]. A 5-PS
of up to 6 lesions at any time during the treatthreshold between 4 and 5 has been proposed as
ment [IR(2)]. Again a biopsy is strongly
positive scan [98]. Moreover, only the associaencouraged: if as expected, the biopsy is negation of 5-PS score 5 in end-of-treatment PET and
tive, the lesion(s) is/are considered negative
high baseline total lesion glycolysis (TLG) in
and should no longer be considered in the folbaseline PET showed a very poor prognosis [98].
lowing scans.
In conclusion, patients with residual FDG uptake
in the context of a nodal mass with bulky disease 3. Increase in FDG uptake of 1 or more lesion(s)
without a concomitant increase in lesion size
at baseline should undergo biopsy if salvage
or number [IR (3)]. Increased immune activtreatment is considered [99].
ity at the site of tumor may manifest as an
increase in FDG uptake. Therefore, by itself,
7.4.5 Treatment Response
changes in uptake should not trigger an assignAssessment to Immune
ment of PD with checkpoint inhibitors.
Checkpoint Inhibitors
Nonetheless, whenever feasible, a biopsy is
advised.
Immune checkpoint inhibitors (CPI) are a new
class of antineoplastic drugs that proved very
Despite an evident bias of subjectivity, a comactive in lung cancer, melanoma, renal carcinoma mon denominator in these three categories of IR
[100–102], and other neoplastic disorders, includ- is a “stable” clinical situation of the patient.
ing HL [20, 103]. In the last few years, the knowl- Moreover, another common recommendation in
edge on the mechanism of action has been largely the follow-up of these three IR is to perform a
unveiled. The activation of the PD-1 receptor on second PET scan 12 weeks after the first imaging
the T-cell surface by its ligand PD-L1 expressed (Fig. 7.4).
on the surface of the neoplastic cells or macrophages downregulates T-cell function, which normally controls the immune activity [104]. Due to 7.4.6 PET in Radiotherapy Planning
this mechanism of action restoring the immune
reactivity of the host against the tumor, the anti- Radiation therapy in HL is an essential therapeuneoplastic activity of CPIs could be delayed, thus tic tool in combined modality treatment (CMT)
fueling a sustained inflammatory response which of early-stage disease and a useful tool to conpersists beyond the drug administration and after solidate chemotherapy treatment in a limited
treatment end. The following criteria and catego- number of advanced-stage patients with a residries for immune response (IRC) in lymphoma ual mass at the end of chemotherapy. Radiation
(LYRIC) have been proposed [105]:
techniques have benefited from an impressive
technological evolution. Extended fields devel1. Increase in overall tumor burden (as assessed oped for single modality treatment have been
by sum of the product of the diameters [SPD]) replaced by more conformal fields designed for
of ≥ 50% of up to 6 measurable lesions in the CMT, encompassing the initially macroscopifirst 12 weeks of therapy, without clinical cally involved tissue volumes in early-stage disdeterioration [IR(1)]. A biopsy is advisable, ease and bulky masses and/or residual masses
but when this is not possible, a second scan after chemotherapy in advanced disease [106–
12 weeks after the initial determination of IR 109]. These changes led to dramatic decreases in
should be planned.
the volume of normal tissue being irradiated and
7
Functional Imaging in Hodgkin Lymphoma
123
a parallel reduction in the risk of serious long-­
term sequelae of RT. A more accurate geometry
for radiation field delineation also demands a
higher accuracy of the imaging technique used.
As FDG-PET has been shown to be more accurate for HL staging, it is by implication also more
precise in defining the initially involved regions
to be irradiated in patients with early-stage disease. No diagnostic modality has a 100% overall
accuracy, and the delineation of the lymphoma
volume must be based on the diagnostic information available, including FDG-PET/CT of both
anatomy and physiology of the disease [110–
112]. In 2013, the International Lymphoma
Radiation Oncology Group (ILROG) identified
the residual gross tumor volume (GTV) as the
first imaging target to be delineated in post-­
chemotherapy FDG-PET to be redrawn in the
­co-­registered images of pre-chemotherapy PET
scan [113]. Then, clinical target volume (CTV)
should be identified by contouring the original
GTV in pre-chemotherapy FDG-PET. Finally,
especially when the target is moving, the internal
target volume (ITV) should be delineated moving
from CTV plus a margin taking into account
uncertainty in size, shape, and position of the
CTV in the patient. The optimal tool to obtain
ITV is using a 4D CT simulator and, as a second
choice, by fluoroscopy. Margins of 1.5–2 cm in
cranio-­caudal direction are commonly chosen in
the chest or upper abdomen. Finally, the planned
target volume (PTV) is calculated to delineate the
IR1: Increase in overall tumor burden (by SPD) of ≥50% of up to 6 measurable lesions in the first 12
weeks of therapy, without clinical deterioration
IR1
Baseline CT
Restaging CT 1-3 wks
Restaging CT 2-7 wks
Restaging CT 3-13 wks
Courtesy H. Jacene
Fig. 7.4 Type 1, type 2, and type 3 IR (indefinite
response) according to LYRIC criteria (From Cheson.
Type 1, type 2 and type 3 LYRIC type of response. Blood
2016 [105]: courtesy of B. Cheson). IR1: Increase in overall tumor burden (by SPD) of ≥50% of up to six measurable lesions in the first 12 weeks of therapy, without
clinical deterioration (Courtesy H. Jacene). IR2:
Appearing of new lesions; or growth of oneor more exist-
ing lesion(s) ≥50%; at any time during treatment; occurring in the context of lack of overall progression (<50% in
increase) of overall tumor burden, by SPD of up to six
lesions at any time during the treatment (Courtesy
H. Jacene). IR3: Increase in FDG uptake of one or more
lesion(s) without a concomitant increase in lesion size or
number (Courtesy L. Schwartz)
A. Gallamini et al.
124
IR2
IR2: Appearance of new lesions; or growth of one or more existing lesion(s) of ≥50%; at any time
during treatment; occurring in the context of lack of overall progression (<50% increase) of overall
tumor burden, by SPD of up to 6 lesions at any time during the treatment.
May 2015
IR2
October 2015
December 2015
Courtesy H. Jacene
Fig. 7.4­ (continued)
7
Functional Imaging in Hodgkin Lymphoma
125
IR3: Increase in FDG uptake of one or more lesion(s) without a concomitant increase in lesion size or number
IR3
L
R
50 pixels
Courtesy L. Schwartz
Fig. 7.4­ (continued)
radiation fields as function of immobilizing ate analysis predictive of progression-free surdevices, body site, and patient cooperation. In vival (PFS) [121]. Different groups [122–124]
early-stage HL undergoing CMT pre-­and a large meta-analysis of the published literachemotherapy, PET/CT should be acquired with ture including also non-Hodgkin lymphoma [125]
the patient in the same position to be used for demonstrated that interim PET (iPET) before
RT. Relatively limited clinical data are available ASCT in r/r HL had a predictive value (PV) indeon CTV definition by FDG-PET for RT planning pendent from other RF. The impact was similar to
in HL [114, 115]. If extended-field irradiation is that during first-line ABVD treatment, where
still used, the impact of FDG-PET is not expected iPET proved the prognostic marker with the highto be very large since additional involvement est PV, irrespective of other RF included in the
found on FDG-PET will often be included in International Prognostic Score (IPS) [15]. More
large treatment fields anyway [116, 117]. With recently, semiquantitative reading of PET with
modern, more conformal radiotherapy, changes metabolic tumor volume (MTV) calculation on
due to FDG-PET have been shown to be signifi- the scan performed at the time of HL relapse
cant [118, 119].
proved to be the strongest predictor of treatment
outcome in r/r HL [126]. Curiously, MTV computed on the FDG-PET performed at relapse and
7.4.7 PET for Response Prediction
interim PET performed immediately before
During Salvage Treatment
ASCT behaved as independent prognostic factors,
and the former improved the PV of the latter.
High-dose chemotherapy (HDC) followed by Overall, the two FDG-PET scans performed in
autologous stem cell transplantation (ASCT) different time points during salvage treatment for
remains the best treatment option for relapsed/ r/r HL were able to recapitulate 4 out of the 5
refractory Hodgkin lymphoma (r/r HL) [120]. RisPACT factors: MTV in baseline PET portrays
More recently, an international consortium mod- tumor burden (bulk and stage IV disease), while
eled r/r HL (the RisPACT consortium) retrospec- iPET is able to detect a low chemosensitivity
tively identified stage IV, ECOG performance (inadequate response to treatment/chemotherapy
status ≥1, bulky nodal lesion >5 cm at relapse, resistance and possibly an inadequate immune
time to relapse ≤3 months, and an inadequate response by the host) [127]. The 3-year EFS rates
therapy response, assessed with a CT or PET/CT for patients with low MTV/negative pretransplant
before ASCT, as the sole risk factors, in multivari- PET, low MTV/positive pretransplant PET, high
A. Gallamini et al.
126
1.0
A. PET neg and low MTV, n=41, censored 38
0.8
B. PET pos and low MTV, n=7, censored 6
65 patients valuable
N=65
BV x 3
BV x 3
Overall 49/65 (75%) of the patients obtained PET-negativity before ASCT
N=18
BV x 3
DS 1-2
PET
PET
MTV
MTV
N=45
ASCT
N=1
DS 3-5
N=31
*auglCE
x2
Lost to f. up N=1
DS 1-2
ASCT
DS 3-5
ASCT
Event free survival
PET-adapted sequential therapy with BV and augICE induces FDG-PET
normalization in 76% of patients with relapsed/ refractory HL
0.6
C. PET neg and high MTV, n=8, censored 3
0.4
0.2
PET
N=47
N=14
*augICE (augmented ICE). :
Ifosfamide 5 gr. /m2 with Mesna 5 gr./m2) day 1
Carboplatin [maximum dose 800 mg]) on day 3,
Etoposide 200 mg/m2 every 12 hours x 3 from day 1; DS: Deauville Score PET; positron emission tomography.
Moskowitz AJ, et at. Blood. 2017;130:2196-203.
0.0
D. PET pos and high MTV, n=3, censored 0
0
10
20
30
40
50
Months
Fig. 7.5 Baseline and interim PET to predict treatment outcome in relapsed/refractory Hodgkin lymphoma (From
Moskowitz AJ Blood 2017, by permission [126])
MTV/negative pretransplant PET, and high MTV/
positive pretransplant PET were 93%, 86%, 38%,
and 0%, respectively (p < 0.0001) [126] (Fig. 7.5).
7.4.8
ET for Follow-Up of HL
P
Patients in Complete
Remission
The impact of surveillance imaging tests (CT,
FDG-PET/CT) during follow-up of lymphoma
patients in CR after treatment is still a matter of
debate. In general, the probability of detecting
relapse during monitoring for disease recurrence
depends on the probability of relapse of the disease in the population being tested, as well as the
sensitivity, specificity, and the frequency of the
test [128]. The prevalence of relapse in a HL
patient in complete remission at the end of treatment is rare, corresponding to one event per 68
visits in HL [129]. Moreover, the risk of relapse
could depend on a number of clinical parameters
associated with disease relapse as follows: (1) the
presence of clinical symptoms, (2) poor chemosensitivity of the tumor at interim evaluation, (3)
the preferred anatomical pattern of recurrence of
a given lymphoma subtype, as well as (4) a residual mass at the end of treatment [130]. Up to 80%
of relapses in HL are associated with clinical
symptoms [129, 131]. Zinzani et al. investigated
the role of surveillance FDG-PET performed
every 6 months for 4 years after CR [132].
Overall, 778 scans were evaluated in a cohort of
160 HL patients. In 11/778 scans (1.4%), PET
was classified as positive, mainly in the first
18 months after CR. All these patients underwent
a confirmatory biopsy: 6/11 were proven positive, mostly in patients having an interim-positive
PET scan. A total of 51 relapses in 160 HL
patients were detected: all 51 by PET/CT, 37 by
CT, and 35 by clinical symptoms. However, 778
PET/CT scans were needed to detect 14 relapses
earlier than CT or other methods. Jerusalem et al.
performed FDG-PET every 4–6 months for 3
years in 36 HL patients who were in CR after
ABVD for a total of 139 scans. Six false-positive
studies, and no false-negative studies, were
found. In the five positive studies, FDG-PET preceded the relapse at a median of 3.5 [1–9] months
[133]. In 2012, El-Galaly et al. published a retrospective study in a cohort of 299 HL patients in
CR after ABVD, in which surveillance PET scan
was used in asymptomatic patients (routine scans
n = 211) or for patients suspected of harboring an
impending relapse (clinically indicated, n = 88).
The true positive rates of routine and clinically
indicated scans were 5% and 13%, respectively.
The overall positive and negative predictive values were 28% and 100%, respectively. The estimated cost per routinely diagnosed relapse was
50,778 US $ [134]. In conclusion, surveillance
FDG-PET cannot be recommended as a routine
follow-up procedure for HL patients in complete
remission after first-line treatment; instead, this
7
Functional Imaging in Hodgkin Lymphoma
should be considered in rare cases of patients
with high risk of relapse without signs or symptoms of disease.
127
chemotherapy only arms were stopped prematurely after a futility interim analysis predicted
failure of meeting the primary endpoint, which
was PFS non-inferiority [136]. The final analyses
confirmed a loss of disease control when radio7.5
PET Response-Adapted
therapy was omitted, even though the difference in
PFS was only clearly clinically relevant in patients
Therapy
without risk factors [7]. In the GHSG HD16 study
7.5.1 Early-Stage HL
for patients with early-stage disease and no risk
factors, patients were randomized to either stanMore than 90% of early-stage HL patients are dard therapy with 2 × ABVD + 20 Gy radiothercured with standard therapy. These patients still apy or an experimental arm where PET-negative
have a reduced life expectancy due to treatment-­ patients after 2 × ABVD would receive no further
related morbidity including second cancers and treatment. After a median follow-up of 47 months,
cardiopulmonary disease. In fact, early-stage HL the estimated 5-year PFS for early PET-negative
patients more often die from late effects of treat- patients was 93.4% (90.4–96.5) with standard
ment than from the disease itself [135]. A number treatment and 86.1% with chemotherapy only. The
of trials have investigated such PET-response-­ hazard ratio was 1.78 with a 95% CI ranging from
adapted therapy in early-stage HL. The UK 1.02 to 3.12, including the non-inferiority margin
National Cancer Research Institute (NCRI) of 3.01. The PFS difference resulted from a signifiLymphoma Group RAPID trial for early-stage cant increase in disease recurrences with infield
patients as well as the German Hodgkin Study recurrence rates of 2.1% vs. 8.7% (p = 0.0003);
Group (GHSG) HD16 randomized trial investi- there was no relevant difference regarding outfield
gated non-inferiority of reducing treatment inten- recurrences (3.7% vs. 4.7%, p = 0.55). There was
sity by omitting radiotherapy interim PET-negative no difference in the estimated 5-year overall surearly-stage patients. The experimental arms of vival in the per-protocol population (98.1% in the
EORTC/GELA/FIL H10 study also omitted radio- standard arm vs. 98.4% in the chemotherapy-­only
therapy in interim PET-negative patients while arm) [137].
escalating from standard ABVD to more intensive
In the H10 study, PET2-positive patients in the
BEACOPPesc followed by radiotherapy in PET- experimental arm (including both favorable and
positive patients. Both the RAPID trial and the unfavorable risk groups) had their treatment intenH10 trial failed to demonstrate non-­inferior PFS in sified with a shift to two cycles of more intensive
the chemotherapy-only arms. In the UK RAPID chemotherapy (BEACOPP escalated) followed by
study, patients who were PET negative after three radiotherapy. In the PET2-positive patients, PFS
cycles of chemotherapy were randomized to either of patients with a BEACOPP-intensified chemoradiotherapy or no further treatment (NFT). At a therapy was significantly better when compared
median follow-up of 3 years, the PFS difference in with the standard arm where patients were treated
the RAPID trial was 3.8% (95% CI, 8.8–1.3) with less intensive chemotherapy followed by
favoring the radiotherapy arm, thus failing to meet radiotherapy (5-year PFS 90.6% vs. 77.4%, HR
the predefined margin of non-inferiority of 7%. 0.42 with 95% CI 0.23–0.74) [7]. This difference
There was no difference in overall survival (OS) was mainly due to improved PFS of patients in the
between the two treatment arms [3]. The European unfavorable risk group who received intensified
H10 trial stratified patients in favorable and unfa- chemotherapy. Even without knowing the results
vorable subgroups, according to standard risk fac- of the German HD16 study, we can conclude from
tors for early-stage HL [7]. All patients had an the RAPID and H10 studies that omitting radioFDG-PET after two cycles of chemotherapy therapy in good-risk (PET2-negative) patients
(PET2). In the experimental arms, PET2-negative results in a modest loss of disease control (PFS),
patients were given chemotherapy alone, but the while there is no difference in OS.
A. Gallamini et al.
128
7.5.2
Advanced-Stage HL
showed similar results from escalation to eBEACOPP in PET-positive patients after two cycles
Around 70% of advanced-stage HL patients can of ABVD (3-year PFS 60%). They also showed
be cured with six cycles of ABVD with or without no improvement of outcomes from consolidative
consolidation radiotherapy, which is the standard radiotherapy to initially bulky masses (>5 cm) in
first-line therapy in most centers. BEACOPPesc patients who were PET2 negative and remained
cures 85–90% of patients if given upfront, but con- posttreatment PET negative [17]. In the German
cern regarding acute toxicity and second neopla- HD18 study, all patients with advanced-stage HL
sias is the reason why a number of centers are had two cycles of BEACOPPesc followed by
reluctant to use this regimen as standard therapy FDG-PET. Patients in the standard arm received
[138]. Numerous trials have investigated PET- eight cycles of BEACOPPesc, but half way durresponse-­adapted therapy for advanced-stage HL ing the study, the duration of standard treatment
patients. Three non-randomized trials used early was abbreviated to six cycles as a result of the
treatment intensification with BEACOPPesc analysis of the German HD15 study. In the exper(The Italian GITIL HD 0607, the US SWOG imental arm, PET2-negative patients (Deauville
S0813 and the UK RATHL trial) or even high- 1–2) only received an additional two cycles of
dose chemotherapy with autologous stem cell sup- BEACOPPesc. The final analysis showed that in
port (The Italian HD0801 FIL trial) in patients PET2-negative patients, a total of four cycles was
who were still PET positive after two cycles of non-inferior to six cycles in terms of 5-year PFS
ABVD. The randomized German HD18 trial (90.8% vs. 92.2%) and importantly was associtested abbreviation on BEACOPPesc therapy ated with significantly less acute and late toxicity
based on PET results after two therapy cycles. The [140]. PET2-positive patients continuing eBEAFrench AHL 2011 trial was also a BEACOPPesc- COPP still had very high 3-year PFS of 87.6%
based randomized trial with treatment modifica- (95% CI 83.0–92.3), illustrating that the positive
tions based on PET after both two and four cycles. predictive value of interim FDG-PET depends on
The UK RATHL study included 1214 patients the treatment setting. A subsequent analysis
with stage IIB–IV or stage IIA with high-risk fea- showed no differences in outcome between
tures who received two cycles of ABVD followed patients in the standard arm who had a Deauville
by FDG-PET. PET2-negative patients (Deauville score of 1–2 and those with a Deauville score of
scores 1–3) were randomized to either continued 3 (the latter would in many other studies be conABVD (total of six cycles) or treatment without sidered PET negative, but in HD18 regarded as
bleomycin (AVD), for 4 cycles. The analyses of PET positive). This indicates that abbreviation of
the PET2-­negative patients showed no difference BEACOPPesc therapy in early PET-negative
in PFS between the two randomization arms patients should be done in those having a PET2
(3-year PFS 85.7% vs. 84.4%) [4]. Even though Deauville score of 3 as well as in patients with
in the primary analysis the results fell short of the Deauville score of 1 and 2 [141]. The French
specified non-inferiority margin, an updated AHL2011 study aimed to assess whether interim
analysis showed that with extended follow-up the FDG-PET could identify good-risk patients to be
study had met its primary endpoint of PFS non-­ given de-escalation treatment BEACOPPesc
inferiority [139].
upfront. A total of 823 patients were randomized
PET2-positive patients received more inten- to the standard arm (n = 413) and the experimensive chemotherapy with an additional four cycles tal arm (n = 410). After a median follow-up of
of BEACOPP14 or BEACOPPesc, resulting in a 50.4 months, the 4-year progression-free survival
3-year PFS of 68%, comparing favorably with (PFS) for the experimental arm was 87.1% and
historical controls from observational studies in for the standard arm 87.4% with no difference in
which ABVD was continued in PET2-positive OS, demonstrating that treatment intensity can be
patients [4, 12, 30]. The Italian single-arm GITIL/ safely reduced from BEACOPPesc to ABVD in
FIL HD0607 (stage IIB–IV patients) study PET2-negative patients [142].
7
Functional Imaging in Hodgkin Lymphoma
Thus, in advanced-stage HL, early interim PET
can be used to de-escalate therapy such as reducing the number of BEACOPPesc cycles, reduction
of intensity from BEACOPPesc to ABVD as well
as omission of bleomycin from ABVD in PET2negative patients. At the same time, intensification
to BEACOPPesc should be considered with insufficient initial PET response to ABVD.
7.5.3
Post-chemotherapy PET/
CT-Driven Consolidation
Radiotherapy
In advanced HL, radiotherapy is used less frequently than in early-stage HL and usually only
to residual masses. CT cannot discriminate
between a residual mass with viable lymphoma
cells and a residual mass consisting only of
fibrotic tissue. However, since PET/CT cannot
detect microscopic disease, it was necessary to
perform studies to investigate whether a PET-­
negative residual mass requires radiotherapy or
not. The mature results of the German HD15 trial
shed light on this for patients treated with
BEACOPP. In the HD15 study, consolidation
radiotherapy was given only to those patients
having a PET-positive single residual mass of
more than 2.5 cm, which was encompassable in a
single field of radiotherapy. The remaining
majority of patients who did not receive radiotherapy had a relapse-free survival of 94% after 1
year, indicating that radiotherapy can be safely
omitted in advanced-stage HL patients who are
PET negative after the end of BEACOPPesc [22,
143]. A retrospective analysis from the British
Columbia Cancer Agency addressed the same
situation for patients treated with ABVD. The
authors reported a 10-year experience with
advanced-stage HL patients having residual
masses >2 cm after chemotherapy undergoing
PET/CT (n = 163). Only patients with a positive
posttreatment PET/CT received radiotherapy. At
the end of treatment, 316 patients had residual
masses larger than 2 cm undergoing PET/CT. Of
those, 264 (83.5%) were FDG-PET negative,
none of whom received RT, and 52 (16.5%) were
FDG-PET positive, of whom 79% (n = 41)
129
received consolidative RT (30–35 Gy). With a
median follow-up for living patients of 4.6 years
(range 0.6–13.5 years), the 5-year freedom from
treatment failure (FFTF) for the whole cohort
was 83%, and the 5-year OS was 94.5%. Not surprisingly, patients with a negative posttreatment
FDG-PET had a superior 5-year FFTF compared
to those with a positive scan (89% vs. 56%,
respectively). In the posttreatment FDG-PET-­
negative group, there was no difference in outcome comparing bulky (n = 112) and non-bulky
(n = 152) subgroups (5-year FFTF 89% vs.
88.5%, respectively). Similarly, when the analysis was restricted to FDG-PET-negative patients
with a bulky mediastinal mass at diagnosis
(n = 102), outcomes were excellent (5-year FFTF
89%; 5-year OS 96%) [84]. These results support
the omission of radiotherapy in advanced-stage
HL patients who achieve a PET-negative remission after six cycles of chemotherapy.
7.5.4
PET/CT-Adapted Therapy
in Relapsed HL
For HL patients in first relapse, a number of risk
factors predict outcome after high-dose chemotherapy with autologous stem cell support
(HD + ASCT) [121]. Two important factors are
the duration of remission prior to relapse and the
response to induction therapy. A number of studies have shown that FDG-PET performed after
induction therapy and before HD + ASCT can
predict the long-term remission after salvage
[144–146]. These studies report a poor long-term
PFS in patients who are FDG-PET positive after
induction chemotherapy (31–41%), compared to
a PFS of 73–82% in patients who reach a PET-­
negative remission before HD + ASCT. A study
from the Memorial Sloan Kettering Cancer Center
investigated a PET-guided approach where FDGPET-positive patients after the standard induction
(ICE, ifosfamide, carboplatin, etoposide) instead
of proceeding to HD + ASCT were given a noncross-resistant regimen consisting of four
biweekly doses of gemcitabine, vinorelbine, and
liposomal
doxorubicin
(GVD)
before
HD + ASCT. Patients who were FDG-­
PET
130
p­ ositive after ICE but became FDG-PET negative
after GVD had similar outcomes compared to
those who were PET negative after ICE [147].
A number of studies demonstrated a high
prognostic value of pretransplant FDG-PET/CT
before alloSCT with reduced intensity [148–
150]. Two studies suggest that FDG-PET/CT
may have a role in guiding the use of donor lymphocyte infusions after allogeneic stem cell
transplantation [151, 152].
A. Gallamini et al.
nitions for complete response (CR) and partial
response (PR), stable disease (SD), progressive
disease (PD), and relapsed disease (RD). Since
these recommendations were based on CT scanning, they included the category of complete
remission unconfirmed (CRu), first introduced in
the Cotswold classification [24], referring to
patients with a persistent mass with size reduction following treatment which, represented
fibrosis, rather than viable lymphoma.
PET scans were initially considered as an
adjunct to lymphoma assessment in about 1990.
7.6
Toward Revised Criteria
Their added value was first demonstrated by
Juweid et al. [153]. They compared the outcome
for PET Scan Interpretation
of 53 patients with diffuse large B-cell lymphoma
Juweid et al. first demonstrated that PET-based using either the NCI-WG response criteria alone
response improved predictability compared to or with the incorporation of PET. Progression-­
criteria relying on CT. They demonstrated that in free survival was similar regardless of whether
patients with diffuse large B-cell NHL and a the patient had a CR or PR based on CT criteria,
residual mass posttreatment, a lack of FDG avid- as long as the mass was not FDG avid. Thus, not
ity correlated closely with a CT-based CR for the only was the accuracy of response assessment
end-of-treatment evaluation [153]. These obser- improved with PET, but the misleading term CRu
vations were subsequently confirmed, leading to was eliminated as a response category.
the incorporation of PET into standardized
Along with the validation of the above obserresponse criteria for patients with DLBCL and vations, numerous other studies demonstrated the
HL [78]. These initial studies of PET used visual superior sensitivity and specificity of PET comassessment using the mediastinal blood pool as pared with CT mandated revised response critethe comparator and formed the basis of the rec- ria. Thus, in 2007 the International Harmonization
ommendations of the International Harmonization Project published recommendations that incorpoProject in lymphoma [78]. Efforts toward a more rated PET [78]. These were primarily for HL and
reliable and reproducible approach resulted in the DLBCL as data with FDG-PET in other histo5-point Deauville criteria which are more accu- logic subtypes were limited.
rate with greater interobserver concordance and a
In the ensuing years following publication of
higher positive predictive value (PPV) [11]. One the 2007 criteria, a large body of data demonconsequence of this improvement was the Lugano strated the role of FDG/PET in follicular lymclassification currently considered the standard phoma (FL) as well as other histologies [155].
for staging and response assessment for all FDG-­ Whereas PET interpretation in the 2007 criteria
avid lymphomas [9, 10].
was purely visual, using mediastinal blood pool
as the reference standard, the newly created
Deauville 5-point scale (D5-PS) used the hepatic
7.6.1 Qualitative vs.
blood pool and was more reproducible [79]. In
Semiquantitative Assessment
2014, the Lugano classification was published
recommending FDG PET/CT as standard for
7.6.1.1 From Anatomical to Functional
staging of FDG-avid histologies [9, 10]. Several
Imaging
studies in HL demonstrated the improved sensiIn 1999, the National Cancer Institute-sponsored tivity of FDG/PET compared with bone marrow
Working Group (NCI-WG) published the first biopsies [31, 52, 57, 156]. Thus, in the Lugano
uniformly adopted response criteria for lym- classification, trephine biopsies were no longer
phoma [154]. These included standardized defi- required in the routine staging of HL. The Lugano
7
Functional Imaging in Hodgkin Lymphoma
classification also recommended the D5PS for
interpretation of FDG/PET for assessment of
response to treatment assessment. The change
from the binary IHP criteria to the five-point DS
is associated with greater interobserver reproducibility and provides greater flexibility to modify
the threshold between a good and a poor response
according to the clinical context and/or the proposed treatment strategy.
The DS is now uniformly adopted and is of
particular importance in risk-adapted therapeutic
strategies. The group that requires particular
attention are the DS5 patients whose outcome
differs significantly even from DS4 [4] in
advanced disease, but has been yet demonstrated
in those with limited-stage disease.
7.6.1.2 T
oward New Criteria
for Response Assessment
of New Drugs
The availability of newer effective treatment led
to issues in interpretation of response and disease
progression due to the occurrence of flare reactions. The most notable examples are immunomodulatory agents that stimulate the immune
system resulting in a flare reaction, particularly
lenalidomide [157–160]. Similar effects were
observed for rituximab [161, 162], brentuximab
vedotin [163], ibrutinib [164], and idelalisib
[165], respectively. In HL, the use of checkpoint
inhibitors can induce a flare reaction that may be
confused with progressive disease [20]. This
potential clinically significant problem led to
publication of the lymphoma response to immunomodulatory therapy (LYRIC) criteria [105]
(see Chap. 4). These recommendations use a provisional term of indeterminate response to categorize various types of flare reactions. Thus,
patients in whom a flare reaction is considered
may remain on treatment until progressive disease is confirmed by biopsy or by continued
growth, or evidence of deterioration.
7.6.1.3 Biomarker Integration
in Response Criteria
Despite the sensitivity of PET-CT scans, the negative predictive value remains imperfect. An
increasingly recognized problem is that relapse
occurs not infrequently, even in patients consid-
131
ered to be in a CMR following therapy. In the
RAPID trial [3], patients with stage I/IIA non-­
bulky disease underwent PET after three cycles
of ABVD. Those with a negative PET were randomized to IFRT or no further therapy. In patients
considered in mCR, there was an apparent correlation between the maximum transverse diameter
of the tumor and the likelihood of recurrence,
with a threshold of 5 cm. Thus, assays that are
able to detect minimal residual disease may be a
useful adjunct, not only in response assessment
but also in distinguishing a flare reaction from
progressive disease. Circulating DNA assays
have been used in patients with NHL to predict
outcome and are approved by the FDA [166].
Only recently has this technology been applied to
HL. Van Eijndhoven et al. [167] suggested that
classical HL-related miRNA levels in circulating
EVs may reflect the presence of tumor tissue and
may provide a marker for therapy response and
relapse monitoring in HL patients. Rossi et al.
[168] using a sensitive next-generation sequencing technique demonstrated the potential for
monitoring patients with HL lymphoma.
In a study of 102 Hodgkin lymphoma patients
treated with ABVD, interim PET, pretreatment
CD68+ cell counts, and the presence of B symptoms were independently associated with PFS
[169]. Therefore, the evaluation of CD68+ cell
counts and B symptoms at diagnosis was suggested to help identify low-risk patients regardless of a positive interim PET result. Agostinelli
et al. [170] provided evidence for an interaction
between the microenvironment and PET. Although
PET after two cycles of therapy predicts outcome
in HL, 10–15% of patients still relapse. When
interim PET after two cycles was combined with
pretreatment environmental markers including
PD-1 and CD68 in Micro Environment cells and
STAT-1 in HRS cells in a Classification and
Regression Tree (CART) analysis, PET2-negative
patients could be distinguished into three risk
groups with markedly different outcomes.
7.6.2
Interim PET
The use of quantitative techniques has been
explored in DLBCL to improve on the accuracy
A. Gallamini et al.
132
of visual assessment. Change in the maximum
SUV (ΔSUVmax) in tumor before and after
treatment has been evaluated as a measure of
response. Lin et al. [171] and Itti et al. [172] performed receiver operator curve (ROC) analysis in
92 patients with DLBCL who underwent PET
scan after two cycles of therapy, and 80 patients
were scanned after four and identified the optimum thresholds for percentage change in
SUVmax for predicting event-free survival
(EFS). The ΔSUVmax better predicted PFS,
compared with standard visual analysis. Biggi
et al. [173] provided data from a retrospective
analysis to suggest that incorporating semiquantitative analyses with visual interpretation
improved predictive value of interim and end-of-­
treatment PET. Unfortunately, despite the fact
that other groups have also reported that
ΔSUVmax predicts response, the thresholds vary
widely among studies. Thus, it is not clear that
this technique is ready for general application
until there is improved consistency in scanning
protocols, matching conditions for serial scans,
as well as proper calibration and scanner maintenance. The optimum cutoff is also likely influenced by timing, with a tendency for a higher
cutoff later during treatment. Clinical correlation
is also critically important.
7.6.3
End-of-Treatment PET
As first reported by Juweid et al. [153] and further recommended in the Lugano classification
[9, 10], PET-CT is the standard of care for remission assessment in FDG-avid patients. PET
allows for better discrimination between fibrosis
and residual tumor, reducing the implementation
of additional therapy. The high-level accuracy for
PET was reported in patients with advanced HL
following ABVD [174]. Another example is the
German HD15 trial in which patients with a
residual mass of at least 2.5 cm underwent PET
[22]. If negative, they received no further therapy.
Patients with a positive scan underwent radiation
therapy. The outcome of the group with a residual
mass that was not FDG avid was similar to those
without a residual mass. Whether the size of the
residual mass is clinically important is controver-
sial. In a subsequent subset analysis of the
German HD 15 trial [175]), those with a positive
end-of-treatment PET and a decrease in the size
of the residual mass <40% had a particularly poor
outcome with a relapse rate in the first year of
23.1% compared with 5.3% for those with a
greater decrease.
For patients with a residual mass following
treatment of HL and for whom additional therapy
is being considered, whether a biopsy is recommended is controversial as it is often falsely negative in HL due to the effects of treatment and
sclerosis of the tumor.
7.7
FDG-PET/MRI
In 2014, a preliminary report was published comparing PET/MRI and PET/CT in a head-to-head
for tumor staging in a cohort of 50 patients
affected by miscellaneous cancers undergoing
PET/CT with a non contrast-enhanced low-dose
CT for attenuation correction at 120 KeV with
10 mA, followed 20 min later by PET-MRI [176].
All patients underwent whole-body PET/CT after
a single intravenous injection of PET tracers 18F-­
FDG, 68Ga-DOTATATE, or 18F-fluoro-ethyl-­
choline (18FFECH) according to a standard
clinical protocol performed on an integrated
64-slice PET/CT scanner (Discovery VCT; GE
Healthcare). PET/MRI imaging was performed
using a Siemens 3T Biograph mMR system with
an integrated PET system within the MR gantry,
which allows simultaneous PET and MR acquisitions without having to reposition the patient.
Two hundred twenty-seven FDG-avid lesions
were found: 225 were detected on PET/CT, and
all the 227 on PET/MRI. Overall, anatomic localization was superior in 5.1% of the cases in PET/
MRI modality compared with PET/CT; this was
attributed to the established superior soft tissue
contrast seen in head and neck, pelvis, and
colorectal cancer patients. More recently,
advances in magnetic resonance imaging (MRI)
technology allowed the introduction of “functional” MRI, such as whole-body diffusion-­
weighted MRI (WB-DW-MRI). The latter could
improve the sensitivity of tumor staging by standard MRI, thanks to a higher tumor-to-­
7
Functional Imaging in Hodgkin Lymphoma
background contrast, and of tumor restaging,
thanks to a different apparent diffusion coefficient of the necrotic post-therapy tissue vs. that of
the viable, pre-therapy, tumor lesion [177].
Herrmann et al. prospectively evaluated the overall accuracy of FDG-PET/CT, FDG-PET/MRI,
and WB-DW-MRI for lymphoma staging and
restaging in a cohort of 61 patients, including 21
cases of HL [178] undergoing sequential scanning with these three imaging techniques [178].
A total of 82 examinations were performed: 14
for staging and 68 for restaging, during (n = 14)
or after chemotherapy (n = 19) and during follow-up (n = 35). Most examinations were performed in HL (n = 28) or diffuse large B-cell
lymphoma (DLBCL; n = 26). FDG-PET/CT was
used as reference method for assessing the performance of the other two techniques. One hundred thirteen lesions were found in lymphoma
staging: FDG-­
PET/MRI detected all of them,
while WB-DW-MRI detected only 64.6% of the
lesions. FDG-PET/CT and FDG-PET/MRI
detected the same number of lesions during
interim (n = 16), end-of-therapy (n = 12), and
surveillance (n = 47) scanning. On the other
hand, WB-DW-MRI detected 56.3%, 50.0%, and
78.7%, respectively. The authors concluded that
FDG-PET/CT and FDG-PET/MRI had a comparable diagnostic performance, while WB-DW-­
MRI could be preferred during patient follow-up
as the latter had a comparable sensitivity to FDGPET/CT or FDG-­PET/MRI in detecting residual,
persisting disease.
7.8
Future Perspectives
7.8.1
Other Tracers
FDG is a glucose analogue, and FDG uptake
reflects the level of glucose metabolism in the tissue. However, like other cancers, lymphoma is
characterized by deregulated cell cycle progression, and most anticancer drugs are designed to
inhibit cell proliferation. Thus, a tracer enabling
imaging of cell proliferation could be useful for
both initial characterization and treatment monitoring of the disease. FDG uptake is somewhat
correlated with cell proliferation, but this correla-
133
tion is weakened by a number of factors, including FDG uptake in nonmalignant lesions as a
result of inflammation or infection. The nucleoside [11C]thymidine was the first PET tracer to
specifically address cell proliferation. Early studies suggested that [11C]thymidine could determine both disease extent and early response to
chemotherapy in aggressive NHL patients [179,
180]. However, the short 20-min half-life of 11C
along with rapid in vivo metabolism limited its
clinical application. The thymidine analogue
3′-deoxy-3′-[18F]fluorothymidine (FLT) offers a
more suitable half-life of 110 min and is stable
in vivo [181]. A pilot study of seven patients
showed that FLT-PET can sensitively identify
lymphoma sites and distinguish aggressive from
indolent NHL [182, 183], although less sensitive
than FDG [184]. Furthermore, preclinical studies
suggested a potential of FLT for imaging in early
response to treatment in lymphoma [185].
Increased uptake of amino acids reflects the
increased transport and protein synthesis of
malignant tissue [186]. This observation provides
the rationale for PET imaging of amino acid
metabolism with the labeled amino acids
L-[methyl-11C]methionine (MET) and O-2-[18F]
fluoroethyl-L-tyrosine (FET). Nuutinen et al.
[187] studied 32 lymphoma patients and found
MET-PET to be highly sensitive for the detection
of disease sites, although there was no correlation
between MET uptake and patient outcome. Kaste
et al. [188] evaluated pediatric patients with lymphoma using 11C-methionine and compared it
directly with FDG PET-CT for staging and follow-­up. Whereas the results were fairly synchronous, 11C-methionine had the disadvantage of
limited abdominal utility because of its uptake in
normal structures. Furthermore, no advantage
has been demonstrated over FDG-PET [189].
Rylova et al. [190] studied immune-PET
imaging
of
CD30-positive
lymphomas
90
using
Zr-desferrioxamine-labeled
CD30-­
specific AC-10 antibody in a preclinical model
and suggested that this technique might be of
value in Hodgkin lymphoma and other CD30-­
expressing tumors. Clinical trials have not been
reported. Thus, at the present time, 18-FDG
remains the preferred tracer for imaging patients
with lymphoma.
A. Gallamini et al.
134
7.8.2
uantitative Methods for PET
Q
Reading (SUV, MTV, TLG)
7.8.2.1 Semiquantitative Assessment
Whereas the DS relies on the qualitative assessment of FDG uptake, Standardized Uptake Value
(SUV), is the most widely used method of PET
interpretation in HL, and more semiquantitative
measures have been explored. The SUV is the
most widely used parameter for quantitative analysis. It is readily available and well established in
routine clinical work. Other investigators evaluated the SUVmax, which is the uptake in the
single voxel exhibiting the highest tracer uptake.
The SUVmax is easily available, has good interreader reproducibility, and is relatively unaffected by partial volume effects. SUVmean is the
mean uptake in a larger user-defined region.
Rossi et al. [191] conducted a retrospective analysis of 59 patients with HL who underwent PET
after two cycles of anthracycline-based chemotherapy. The DS was compared with the ΔSUV
comparing PET-0 with PET-2. A good response
was considered a ΔSUV of at least 71%. Using
the DS, 78% were considered to have a negative
scan, seven of whom failed treatment for a NPV
of 85%. The ΔSUVmax was greater than 71% in
83% of patients, six of whom failed treatment
with a NPV of 88%. On the other hand, the PPV
was superior for the ΔSUVmax at 70% vs. 46%
using the DS. The ΔSUVmax reclassified 46% of
the DS-positive patients as negative. Moreover,
the ΔSUVmax was more accurate at predicting
4-year PFS 82% vs. 30%. Unfortunately, the
ΔSUV is not yet ready to be adopted in general
practice. This method of interpretation is not
standardized, with different cutoffs in the various
published studies. Further studies should clarify
the value compared with standard interpretation.
7.8.2.2 Metabolic Tumor Volume
A number of prognostic scoring systems are
used by various investigators in patients with
HL. Most of these include factors that are surrogates for tumor burden. The first attempt to
approach the total tumor burden (TB) was made
by Specht et al. [192] evaluating 290 patients
with limited-­stage HL. Data were supplied by
the Danish National Hodgkin Study, with an
index combining the tumor size of each involved
region with the number of involved regions, sex,
and histologic subtype to identify patients at a
particularly poor risk following radiotherapy or
chemoradiotherapy. Subsequently, other systems have been proposed using functional imaging. PET-CT affords the opportunity to provide
a better assessment of tumor burden by measuring the total metabolic tumor volume (TMTV),
which is related to both the tumor size and the
activity of tumor and microenvironment cells.
For these reasons, baseline TMTV has been
suggested by several investigators to be a new
risk factor to stratify early-stage patients. A high
TMTV at baseline predicts a lower survival in
early-stage HL [193, 194]. Using PET-CT,
Cottereau et al. [60] evaluated the TMTV in 258
evaluable, high-risk, early-stage HL patients,
treated on the H10 trial, 101 with favorable and
157 unfavorable disease. TMTV accurately predicted PFS and OS with 86% and 84% specificity, respectively. Indeed, TMTV identified 4 risk
groups and was more accurate than baseline
staging systems from the European Organization
for the Research and Treatment of Cancer
(EORTC), the German Hodgkin Study Group,
the NCCN Network, and the Groupe d’Etude
des Lymphomes de l’Adulte. The German
Hodgkin Study Group evaluated 310 patients
who underwent PET for staging and calculated
MTV by four different methods [195] and found
all predicted outcomes.
Moskowitz [126] et al. reported a phase II,
risk-adapted study in patients with relapsed or
refractory HL. Transplant eligible patients
received two or three cycles of brentuximab
vedotin; those who became PET negative went
on to ASCT. Those with residual disease received
augmented ifosfamide, carboplatin, and etoposide (ICE) prior to ASCT. Three-year overall survival and event-free survival (EFS) were 95%
and 82%, respectively. Factors predicting favorable outcome were low baseline metabolic tumor
volume (<109.5 cm3) and relapsed disease, which
was associated with a 100% likelihood of 3-year
EFS. For patients who underwent ASCT, bMTV
and pre-ASCT PET were independently prognos-
7
Functional Imaging in Hodgkin Lymphoma
tic. The 3-year EFS for pre-ASCT PET-positive
patients with low bMTV was 86%. Thus, bMTV
improved the predictive power of pre-ASCT
PET. In the future, consideration for ASCT
should include such factors, to treat those patients
most likely to benefit and to exclude those
unlikely to do so.
The use of interim PET remains controversial.
It appears to permit escalation of treatment
toward improved outcome in PET-positive
patients HL, as well as a reduction in treatment
for those whose scan is interpreted as negative.
Although PET-2 is the most common time point
for an interim study, other data suggest that PET
as early as after the first cycle may be preferred.
Hutchings et al. found a very high prognostic
value of PET after one cycle of chemotherapy
and a higher negative predictive value after one
cycle than after two cycles of chemotherapy [66].
The authors concluded that PET after one cycle
should be the preferred method for PET-response-­
adapted strategies designed to select patients candidate to a less intensive or a de-escalated
treatment. Kostakoglu et al. has come to a similar
conclusion [14]. However, earlier studies may be
associated with a higher false positive rate [14].
Thus, at the present time, two cycles is the current standard.
7.9
General Recommendations
for the Use of PET in HL
Cheson et al. [9] recommended PET-CT as part
of routine staging and restaging of patients with
HL in the Lugano classification. A contrast-­
enhanced CT is not required unless measurement
of a mass is critical to developing a treatment
strategy. When used for initial staging, ample
data support that a bone marrow aspirate and
biopsy are no longer required. Interim PET may
also be a valuable adjunct in Hodgkin lymphoma,
as described above. A restaging study is best performed 6–8 weeks following the completion of
standard induction therapy, with no contrast-­
enhanced CT indicated as whether there is evidence of persistent disease or not is what is
critical. No additional scans are needed for sur-
135
veillance unless there is clinical evidence supporting disease recurrence. PET-CT provides
important guidance for patients who relapse and
are being considered candidates for intensive
therapy with a stem cell transplant.
Overall, PET-CT plays a major role in risk-­
adapted approaches as part of initial treatment,
during interim evaluation, and in conjunction
with other factors for patients being considered
for various salvage strategies.
References
1. Connors JM, Jurczak W, Straus DJ, Ansell SM, Kim
WS, Gallamini A et al (2018) Brentuximab vedotin
with chemotherapy for stage III or IV Hodgkin’s lymphoma. N Engl J Med 378(4):331–344
2. Diehl V, Franklin J, Pfreundschuh M, Lathan B,
Paulus U, Hasenclever D et al (2003) Standard and
increased-dose BEACOPP chemotherapy compared
with COPP-ABVD for advanced Hodgkin's disease.
N Engl J Med 348(24):2386–2395
3. Radford J, Illidge T, Counsell N, Hancock B,
Pettengell R, Johnson P et al (2015) Results of a trial
of PET-directed therapy for early-stage Hodgkin’s
lymphoma. N Engl J Med 372(17):1598–1607
4. Johnson P, Federico M, Kirkwood A, Fossa A, Berkahn
L, Carella A et al (2016) Adapted treatment guided
by interim PET-CT scan in advanced Hodgkin’s lymphoma. N Engl J Med 374(25):2419–2429
5. Press OW, Li H, Schoder H, Straus DJ, Moskowitz
CH, LeBlanc M et al (2016) US intergroup trial of
response-adapted therapy for stage III to IV Hodgkin
lymphoma using early interim Fluorodeoxyglucose-­
positron emission tomography imaging: southwest oncology group S0816. J Clin Oncol
34(17):2020–2027
6. Zinzani PL, Broccoli A, Gioia DM, Castagnoli A,
Ciccone G, Evangelista A et al (2016) Interim positron emission tomography response-adapted therapy
in advanced-stage Hodgkin lymphoma: final results of
the phase II part of the HD0801 study. J Clin Oncol
34(12):1376–1385
7. Andre MPE, Girinsky T, Federico M, Reman O,
Fortpied C, Gotti M et al (2017) Early positron emission tomography response-adapted treatment in stage
I and II Hodgkin lymphoma: final results of the randomized EORTC/LYSA/FIL H10 trial. J Clin Oncol
35(16):1786–1794
8. Cheson BD (2011) Role of functional imaging
in the management of lymphoma. J Clin Oncol
29(14):1844–1854
9. Cheson BD, Fisher RI, Barrington SF, Cavalli F,
Schwartz LH, Zucca E et al (2014) Recommendations
for initial evaluation, staging, and response assessment
A. Gallamini et al.
136
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
of Hodgkin and non-Hodgkin lymphoma: the Lugano
classification. J Clin Oncol 32(27):3059–3068
Barrington SF, Mikhaeel NG, Kostakoglu L, Meignan
M, Hutchings M, Mueller SP et al (2014) Role of
imaging in the staging and response assessment of
lymphoma: consensus of the international conference
on malignant lymphomas imaging working group. J
Clin Oncol 32(27):3048–3058
Biggi A, Gallamini A, Chauvie S, Hutchings M,
Kostakoglu L, Gregianin M et al (2013) International
validation study for interim PET in ABVD-treated,
advanced-stage Hodgkin lymphoma: interpretation
criteria and concordance rate among reviewers. J Nucl
Med 54(5):683–690
Gallamini A, Barrington SF, Biggi A, Chauvie S,
Kostakoglu L, Gregianin M et al (2014) The predictive role of interim positron emission tomography for
Hodgkin lymphoma treatment outcome is confirmed
using the interpretation criteria of the Deauville five-­
point scale. Haematologica 99(6):1107–1113
Hutchings M, Loft A, Hansen M, Pedersen LM,
Buhl T, Jurlander J et al (2006) FDG-PET after two
cycles of chemotherapy predicts treatment failure
and progression-­free survival in Hodgkin lymphoma.
Blood 107(1):52–59
Kostakoglu L, Goldsmith SJ, Leonard JP, Christos P,
Furman RR, Atasever T et al (2006) FDG-PET after
1 cycle of therapy predicts outcome in diffuse large
cell lymphoma and classic Hodgkin disease. Cancer
107(11):2678–2687
Gallamini A, Hutchings M, Rigacci L, Specht L, Merli
F, Hansen M et al (2007) Early interim 2-[18F]fluoro2-deoxy-D-glucose positron emission tomography
is prognostically superior to international prognostic score in advanced-stage Hodgkin’s lymphoma: a
report from a joint Italian-Danish study. J Clin Oncol
25(24):3746–3752
Rigacci L, Puccini B, Zinzani PL, Biggi A,
Castagnoli A, Merli F et al (2015) The prognostic
value of positron emission tomography performed
after two courses (INTERIM-PET) of standard therapy on treatment outcome in early stage Hodgkin
lymphoma: a multicentric study by the fondazione italiana linfomi (FIL). Am J Hematol 90(6):
499–503
Gallamini A, Tarella C, Viviani S, Rossi A, Patti C,
Mule A et al (2018) Early chemotherapy intensification with escalated BEACOPP in patients with
advanced-stage Hodgkin lymphoma with a positive
interim positron emission tomography/computed
tomography scan after two ABVD cycles: long-term
results of the GITIL/FIL HD 0607 trial. J Clin Oncol
36(5):454–462
Roemer MG, Advani RH, Ligon AH, Natkunam Y,
Redd RA, Homer H et al (2016) PD-L1 and PD-L2
genetic alterations define classical Hodgkin lymphoma
and predict outcome. J Clin Oncol 34(23):2690–2697
Ansell SM, Lesokhin AM, Borrello I, Halwani A,
Scott EC, Gutierrez M et al (2015) PD-1 blockade
with nivolumab in relapsed or refractory Hodgkin’s
lymphoma. N Engl J Med 372(4):311–319
20. Armand P, Engert A, Younes A, Fanale M, Santoro
A, Zinzani PL et al (2018) Nivolumab for relapsed/
refractory classic Hodgkin lymphoma after failure
of autologous hematopoietic cell transplantation:
extended follow-up of the multicohort single-­
arm
phase II CheckMate 205 trial. J Clin Oncol 36(14):
1428–1439
21. Chen R, Zinzani PL, Fanale MA, Armand P, Johnson
NA, Brice P et al (2017) Phase II study of the efficacy and safety of Pembrolizumab for relapsed/
refractory classic Hodgkin lymphoma. J Clin Oncol
35(19):2125–2132
22. Engert A, Haverkamp H, Kobe C, Markova J, Renner
C, Ho A et al (2012) Reduced-intensity chemotherapy and PET-guided radiotherapy in patients with
advanced-stage Hodgkin’s lymphoma (HD15 trial): a
randomised, open-label, phase 3 non-inferiority trial.
Lancet 379(9828):1791–1799
23. Carbone PP, Kaplan HS, Musshoff K, Smithers
DW, Tubiana M (1971) Report of the committee on
Hodgkin’s disease staging classification. Cancer Res
31(11):1860–1861
24. Lister TA, Crowther D, Sutcliffe SB, Glatstein E,
Canellos GP, Young RC et al (1989) Report of a committee convened to discuss the evaluation and staging
of patients with Hodgkin’s disease: Cotswolds meeting. J Clin Oncol 7(11):1630–1636
25. Glatstein E, Guernsey JM, Rosenberg SA, Kaplan
HS (1969) The value of laparotomy and splenectomy in the staging of Hodgkin’s disease. Cancer
24(4):709–718
26. Castellino RA, Hoppe RT, Blank N, Young SW,
Neumann C, Rosenberg SA et al (1984) Computed
tomography, lymphography, and staging laparotomy:
correlations in initial staging of Hodgkin disease. AJR
Am J Roentgenol 143(1):37–41
27. DeVita VT Jr, Simon RM, Hubbard SM, Young RC,
Berard CW, Moxley JH III et al (1980) Curability
of advanced Hodgkin’s disease with chemotherapy.
Long-term follow-up of MOPP-treated patients
at the National Cancer Institute. Ann Intern Med
92(5):587–595
28. Gospodarowicz MK, O'Sullivan B, Koh ES (2006)
Prognostic factors: principles and applications. In:
Gospodarowicz MK, O’Sullivan B, Sobin LH (eds)
Prognostic factors in cancer, 3rd edn. Wiley-Liss,
Hoboken, NJ, pp 23–28
29. Diehl V, Stein H, Hummel M, Zollinger R, Connors
JM (2003) Hodgkin’s lymphoma: biology and treatment strategies for primary, refractory, and relapsed
disease. Hematology Am Soc Hematol Educ Program
2003:225–247
30. Hutchings M, Loft A, Hansen M, Pedersen LM,
Berthelsen AK, Keiding S et al (2006) Position emission tomography with or without computed tomography in the primary staging of Hodgkin’s lymphoma.
Haematologica 91(4):482–489
31. El Galaly TC, d'Amore F, Mylam KJ, Nully Brown
P, Bgsted M, Bukh A et al (2012) Routine bone
marrow biopsy has little or no therapeutic consequence for positron emission tomography/computed
7
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
Functional Imaging in Hodgkin Lymphoma
tomography-­
staged treatment-naive patients with
Hodgkin lymphoma. J Clin Oncol 30(36):4508–4514
Alzahrani M, El-Galaly TC, Hutchings M, Hansen
JW, Loft A, Johnsen HE et al (2016) The value of
routine bone marrow biopsy in patients with diffuse
large B-cell lymphoma staged with PET/CT: a DanishCanadian study. Ann Oncol 27(6):1095–1099
Thompson CJ (2002) Instrumentation. In: Wahl
RL (ed) Principles and practice of positron emission tomography. Lippincott Williams & Wilkins,
Philadelphia, PA, pp 48–64
Finn RD, Schlyer DJ (2002) Production of radionuclides for PET. In: Wahl RL (ed) Principles and
practice of positron emission tomography. Lippincott
Williams & Wilkins, Philadelphia, PA, pp 1–15
Fowler JS, Ding Y (2002) Chemistry. In: Wahl RL (ed)
Principles and practice of positron emission tomography. Lippincott Williams & Wilkins, Philadelphia,
PA, pp 16–47
Ell PJ, von Schulthess GKPET (2002) CT: a new road
map. Eur J Nucl Med Mol Imaging 29(6):719–720
Warburg O (1926) Über den Stoffwechsel der
Tumoren: arbeiten aus dem Kaiser Wilhelm-Institut
für Biologie, Berlin-Dahlem. Springer, Berlin
Au KK, Liong E, Li JY, Li PS, Liew CC, Kwok TT
et al (1997) Increases in mRNA levels of glucose
transporters types 1 and 3 in Ehrlich ascites tumor
cells during tumor development. J Cell Biochem
67(1):131–135
Yamamoto T, Seino Y, Fukumoto H, Koh G, Yano H,
Inagaki N et al (1990) Over-expression of facilitative
glucose transporter genes in human cancer. Biochem
Biophys Res Commun 170(1):223–230
Brown RS, Wahl RL (1993) Overexpression of Glut-1
glucose transporter in human breast cancer. An immunohistochemical study. Cancer 72(10):2979–2985
Aloj L, Caraco C, Jagoda E, Eckelman WC, Neumann
RD (1999) Glut-1 and hexokinase expression: relationship with 2-fluoro-2-deoxy-D-glucose uptake
in A431 and T47D cells in culture. Cancer Res
59(18):4709–4714
Higashi K, Clavo AC, Wahl RL (1993) Does FDG
uptake measure proliferative activity of human cancer cells? In vitro comparison with DNA flow cytometry and tritiated thymidine uptake. J Nucl Med
34(3):414–419
Brown RS, Leung JY, Fisher SJ, Frey KA, Ethier SP,
Wahl RL (1996) Intratumoral distribution of tritiated-­
FDG in breast carcinoma: correlation between
Glut-1 expression and FDG uptake. J Nucl Med
37(6):1042–1047
Wahl RL, Henry CA, Ethier SP (1992) Serum glucose: effects on tumor and normal tissue accumulation
of 2-[F-18]-fluoro-2-deoxy-D-glucose in rodents with
mammary carcinoma. Radiology 183(3):643–647
Clavo AC, Brown RS, Wahl RL (1995)
Fluorodeoxyglucose uptake in human cancer cell lines
is increased by hypoxia. J Nucl Med 36(9):1625–1632
Kubota R, Yamada S, Kubota K, Ishiwata K,
Tamahashi N, Ido T (1992) Intratumoral distribution of fluorine-18-fluorodeoxyglucose in vivo: high
137
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
accumulation in macrophages and granulation tissues studied by microautoradiography. J Nucl Med
33(11):1972–1980
Brown RS, Leung JY, Fisher SJ, Frey KA, Ethier
SP, Wahl RL (1995) Intratumoral distribution of
tritiated fluorodeoxyglucose in breast carcinoma:
I. Are inflammatory cells important? J Nucl Med
36(10):1854–1861
Higashi K, Clavo AC, Wahl RL (1993) In vitro assessment of 2-fluoro-2-deoxy-D-glucose, L-methionine
and thymidine as agents to monitor the early response
of a human adenocarcinoma cell line to radiotherapy.
J Nucl Med 34(5):773–779
Spaepen K, Stroobants S, Dupont P, Bormans G,
Balzarini J, Verhoef G et al (2003) [(18)F]FDG PET
monitoring of tumour response to chemotherapy:
does [(18)F]FDG uptake correlate with the viable
tumour cell fraction? Eur J Nucl Med Mol Imaging
30(5):682–688
Hutchings M, Loft A, El-Galaly TC (2016) PET/
CT for Hodgkin lymphoma staging. In: Gallamini
A (ed) PET scan in Hodgkin lymphoma. Springer,
Heidelberg, pp 1–13
Bednaruk-Mlynski E, Pienkowska JF, Skorzak
AF, Malkowski BF, Kulikowski WF, Subocz EF,
Dzietczenia J et al (2015) Comparison of positron
emission tomography/computed tomography with
classical contrast-enhanced computed tomography
in the initial staging of Hodgkin lymphoma. Leuk
Lymphoma 56(2):377–382
Barrington SF, Kirkwood AA, Franceschetto A,
Fulham MJ, Roberts TH, Almquist H et al (2016)
PET-CT for staging and early response: results from
the response-adapted therapy in advanced Hodgkin
lymphoma study. Blood 127(12):1531–1538
Weiler-Sagie M, Kagna O, Dann EJ, Ben-Barak A,
Israel O (2014) Characterizing bone marrow involvement in Hodgkin’s lymphoma by FDG-PET/CT. Eur J
Nucl Med Mol Imaging 41(6):1133–1140
Zwarthoed C, El-Galaly TC, Canepari M, Ouvrier
MJ, Viotti J, Ettaiche M et al (2017) Prognostic
value of bone marrow tracer uptake pattern in baseline PET scans in Hodgkin lymphoma: results from
an international collaborative study. J Nucl Med
58(8):1249–1254
Levis A, Pietrasanta D, Godio L, Vitolo U, Ciravegna
G, Di Vito F et al (2004) A large-scale study of bone
marrow involvement in patients with Hodgkin’s lymphoma. Clin Lymphoma 5(1):50–55
Brunning RD, Bloomfield CD, McKenna RW,
Peterson LA (1975) Bilateral trephine bone marrow
biopsies in lymphoma and other neoplastic diseases.
Ann Intern Med 82(3):365–366
Voltin CA, Goergen H, Baues C, Fuchs M, Mettler
J, Kreissl S et al (2018) Value of bone marrow
biopsy in Hodgkin lymphoma patients staged by
FDG PET: results from the German Hodgkin study
group trials HD16, HD17, and HD18. Ann Oncol
29(9):1926–1931
Purz S, Mauz-Korholz C, Korholz D, Hasenclever
D, Krausse A, Sorge I et al (2011) [18F]
A. Gallamini et al.
138
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
Fluorodeoxyglucose positron emission tomography
for detection of bone marrow involvement in children
and adolescents with Hodgkin’s lymphoma. J Clin
Oncol 29(26):3523–3528
Ferme C, Eghbali H, Meerwaldt JH, Rieux C, Bosq
J, Berger F et al (2007) Chemotherapy plus involved-­
field radiation in early-stage Hodgkin’s disease. N
Engl J Med 357(19):1916–1927
Cottereau AS, Versari A, Loft A, Casasnovas O, Bellei
M, Ricci R et al (2018) Prognostic value of baseline
metabolic tumor volume in early-stage Hodgkin lymphoma in the standard arm of the H10 trial. Blood
131(13):1456–1463
Kostakoglu L, Chauvie S (2018) Metabolic tumor
volume metrics in lymphoma. Semin Nucl Med
48(1):50–66
Therasse P, Arbuck SG, Eisenhauer EA, Wanders J,
Kaplan RS, Rubinstein L et al (2000) New guidelines
to evaluate the response to treatment in solid tumors.
European Organization for Research and Treatment
of Cancer, National Cancer Institute of the United
States, National Cancer Institute of Canada. J Natl
Cancer Inst 92(3):205–216
Radford JA, Cowan RA, Flanagan M, Dunn G,
Crowther D, Johnson RJ et al (1988) The significance
of residual mediastinal abnormality on the chest
radiograph following treatment for Hodgkin’s disease. J Clin Oncol 6(6):940–946
Surbone A, Longo DL, DeVita VT Jr, Ihde DC,
Duffey PL, Jaffe ES et al (1988) Residual abdominal
masses in aggressive non-Hodgkin’s lymphoma after
combination chemotherapy: significance and management. J Clin Oncol 6(12):1832–1837
Naumann R, Vaic A, Beuthien-Baumann B, Bredow
J, Kropp J, Kittner T et al (2001) Prognostic value
of positron emission tomography in the evaluation of post-treatment residual mass in patients with
Hodgkin's disease and non-Hodgkin’s lymphoma. Br
J Haematol 115(4):793–800
Hutchings M, Kostakoglu L, Zaucha JM, Malkowski
B, Biggi A, Danielewicz I et al (2014) In vivo treatment sensitivity testing with positron emission
tomography/computed tomography after one cycle of
chemotherapy for Hodgkin lymphoma. J Clin Oncol
32(25):2705–2711
Zaucha JM, Malkowski B, Chauvie S, Subocz E,
Tajer J, Kulikowski W et al (2017) The predictive
role of interim PET after the first chemotherapy cycle
and sequential evaluation of response to ABVD in
Hodgkin’s lymphoma patients-the polish lymphoma
research group (PLRG) observational study. Ann
Oncol 28(12):3051–3057
Cerci JJ, Pracchia LF, Linardi CC, Pitella FA,
Delbeke D, Izaki M et al (2010) 18F-FDG PET after
2 cycles of ABVD predicts event-free survival in
early and advanced Hodgkin lymphoma. J Nucl Med
51(9):1337–1343
Zinzani P, Rigacci L, Stefoni V, Broccoli A, Puccini
B, Castagnoli A et al (2012) Early interim 18FFDG PET in Hodgkin’s lymphoma: evaluation on
70.
71.
72.
73.
74.
75.
76.
77.
78.
79.
80.
81.
304 patients. Eur J Nucl Med Mol Imaging 39(1):
4–12
Terasawa T, Lau J, Bardet S, Couturier O,
Hotta T, Hutchings M et al (2009) Fluorine-18-­
fluorodeoxyglucose positron emission tomography
for interim response assessment of advanced-stage
Hodgkin’s lymphoma and diffuse large B-cell
lymphoma: a systematic review. J Clin Oncol
27(11):1906–1914
Maynard J, Emmas SA, Ble FX, Barjat H,
Lawrie E, Hancox U et al (2016) The use of (18)
F-fluorodeoxyglucose positron emission tomography
((18)F-FDG PET) as a pathway-specific biomarker
with AZD8186, a PI3Kbeta/delta inhibitor. EJNMMI
Res 6(1):62
Simontacchi G, Filippi AR, Ciammella P, Buglione
M, Saieva C, Magrini SM et al (2015) Interim PET
after two ABVD cycles in early-stage Hodgkin lymphoma: outcomes following the continuation of chemotherapy plus radiotherapy. Int J Radiat Oncol Biol
Phys 92(5):1077–1083
Evens AM, Kostakoglu L (2014) The role of FDG-­
PET in defining prognosis of Hodgkin lymphoma for
early-stage disease. Blood 124(23):3356–3364
Sher DJ, Mauch PM, Van Den Abbeele A, LaCasce
AS, Czerminski J, Ng AK (2009) Prognostic significance of mid- and post-ABVD PET imaging in
Hodgkin’s lymphoma: the importance of involved-­
field radiotherapy. Ann Oncol 20(11):1848–1853
Engles JM, Quarless SA, Mambo E, Ishimori T,
Cho SY, Wahl RL (2006) Stunning and its effect on
3H-FDG uptake and key gene expression in breast
cancer cells undergoing chemotherapy. J Nucl Med
47(4):603–608
Kasamon YL, Jones RJ, Wahl RL (2007) Integrating
PET and PET/CT into the risk-adapted therapy of
lymphoma. J Nucl Med 48(Suppl 1):19S–27S
Zijlstra JM, Lindauer-van der Werf G, Hoekstra OS,
Hooft L, Riphagen II, Huijgens PC (2006) 18F-fluorodeoxyglucose positron emission tomography for
post-treatment evaluation of malignant lymphoma: a
systematic review. Haematologica 91(4):522–529
Cheson BD, Pfistner B, Juweid ME, Gascoyne RD,
Specht L, Horning SJ et al (2007) Revised response
criteria for malignant lymphoma. J Clin Oncol
25(5):579–586
Meignan M, Gallamini A, Meignan M, Gallamini
A, Haioun C (2009) Report on the first international
workshop on interim-PET scan in lymphoma. Leuk
Lymphoma 50(8):1257–1260
Barnes JA, LaCasce AS, Zukotynski K, Israel D,
Feng Y, Neuberg D et al (2011) End-of-treatment
but not interim PET scan predicts outcome in nonbulky limited-stage Hodgkin’s lymphoma. Ann Oncol
22(4):910–915
Fallanca F, Alongi P, Incerti E, Gianolli L, Picchio
M, Kayani I et al (2016) Diagnostic accuracy of FDG
PET/CT for clinical evaluation at the end of treatment
of HL and NHL: a comparison of the Deauville criteria (DC) and the international harmonization proj-
7
82.
83.
84.
85.
86.
87.
88.
89.
90.
91.
92.
Functional Imaging in Hodgkin Lymphoma
ect criteria (IHPC). Eur J Nucl Med Mol Imaging
43(10):1837–1848
Jerusalem G, Beguin Y, Fassotte MF, Najjar F, Paulus
P, Rigo P et al (1999) Whole-body positron emission tomography using 18F-fluorodeoxyglucose for
posttreatment evaluation in Hodgkin's disease and
non-Hodgkin's lymphoma has higher diagnostic and
prognostic value than classical computed tomography
scan imaging. Blood 94(2):429–433
Terasawa T, Nihashi T, Hotta T, Nagai H (2008) 18F-­
FDG PET for posttherapy assessment of Hodgkin’s
disease and aggressive non-Hodgkin's lymphoma: a
systematic review. J Nucl Med 49(1):13–21
Savage KJ, Connors JM, Villa DR, Hapgood G,
Gerrie AS, Shenkier TN et al (2015) Advanced-stage
classical Hodgkin lymphoma patients with a negative PET-scan following treatment with ABVD have
excellent outcomes without the need for consolidative
radiotherapy regardless of disease bulk at presentation. Blood 126(23):579
Picardi M, De Renzo A, Pane F, Nicolai E, Pacelli
R, Salvatore M et al (2007) Randomized comparison of consolidation radiation versus observation in
bulky Hodgkin’s lymphoma with post-chemotherapy
negative positron emission tomography scans. Leuk
Lymphoma 48(9):1721–1727
Hutchings M, Mikhaeel NG, Fields PA, Nunan T,
Timothy AR (2005) Prognostic value of interim
FDG-PET after two or three cycles of chemotherapy in Hodgkin lymphoma. Ann Oncol 16(7):
1160–1168
Gallamini A, Fiore F, Sorasio R, Meignan M
(2009) Interim positron emission tomography scan
in Hodgkin lymphoma: definitions, interpretation rules, and clinical validation. Leuk Lymphoma
50(11):1761–1764
Hoppe RT, Advani RH, Ai WZ, Ambinder RF, Aoun
P, Armand P et al (2018) NCCN guidelines insights:
Hodgkin lymphoma, version 1. J Natl Compr Canc
Netw 16(3):245–254
Eichenauer DA, Aleman BMP, Andre M, Federico
M, Hutchings M, Illidge T et al (2018) Hodgkin lymphoma: ESMO clinical practice guidelines for diagnosis, treatment and follow-up. Ann Oncol 29:iv19
Follows GA, Ardeshna KM, Barrington SF, Culligan
DJ, Hoskin PJ, Linch D et al (2014) Guidelines for
the first line management of classical Hodgkin lymphoma. Br J Haematol 166(1):34–49
Barrington SF, Qian W, Somer EJ, Franceschetto
A, Bagni B, Brun E et al (2010) Concordance
between four European centres of PET reporting criteria designed for use in multicentre trials in
Hodgkin lymphoma. Eur J Nucl Med Mol Imaging
37(10):1824–1833
Hasenclever D, Kurch LF, Mauz-Korholz CF,
Elsner AF, Georgi TF, Wallace HF et al (2014)
qPET – a quantitative extension of the Deauville
scale to assess response in interim FDG-PET scans
in lymphoma. Eur J Nucl Med Mol Imaging 41(7):
1301–1308
139
93. Canellos GP (1988) Residual mass in lymphoma
may not be residual disease. J Clin Oncol 6(6):
931–933
94. van BK, Kelta M, Bahaguna P (2001) Primary mediastinal B-cell lymphoma: a review of pathology and
management. J Clin Oncol 19(6):1855–1864
95. Savage KJ, Monti S, Kutok JL, Cattoretti G, Neuberg
D, De LL et al (2003) The molecular signature of
mediastinal large B-cell lymphoma differs from that
of other diffuse large B-cell lymphomas and shares
features with classical Hodgkin lymphoma. Blood
102(12):3871–3879
96. Martelli M, Ceriani L, Zucca E, Zinzani PL, Ferreri
AJ, Vitolo U et al (2014) [18F]fluorodeoxyglucose
positron emission tomography predicts survival after
chemoimmunotherapy for primary mediastinal large
B-cell lymphoma: results of the international extranodal lymphoma study group IELSG-26 study. J
Clin Oncol 32(17):1769–1775
97. Melani C, Advani R, Roschewski M, Walters KM,
Chen CC, Baratto L et al (2018) End-of-treatment
and serial PET imaging in primary mediastinal B-cell lymphoma following dose-adjusted
EPOCH-R: a paradigm shift in clinical decision
making. Haematologica 103(8):1337–1344
98. Pinnix CC, Ng AK, Dabaja BS, Milgrom SA,
Gunther JR, Fuller CD et al (2018) Positron emission
tomography-computed tomography predictors of
progression after DA-R-EPOCH for PMBCL. Blood
Adv 2(11):1334–1343
99. Barrington SF, Kluge R (2017) FDG PET for therapy monitoring in Hodgkin and non-Hodgkin lymphomas. Eur J Nucl Med Mol Imaging 44(Suppl
1):97–110
100. Robert C, Schachter JF, Long GV, Arance A, Arance
AF, Grob JJ, Mortier L, Mortier LF et al (2015)
Pembrolizumab versus ipilimumab in advanced
melanoma. N Engl J Med 372(26):2521–2532
101. Herbst RS, Baas P, Kim DW, Felip E, Perez-­
Gracia JL, Han JY et al (2016) Pembrolizumab
versus docetaxel for previously treated, PD-L1-­
positive, advanced non-small-cell lung cancer
(KEYNOTE-010): a randomised controlled trial.
Lancet 387(10027):1540–1550
102. Motzer RJ, Escudier B, McDermott DF, George S,
Hammers HJ, Srinivas S et al (2015) Nivolumab versus everolimus in advanced renal-cell carcinoma. N
Engl J Med 373(19):1803–1813
103. Armand P, Shipp MA, Ribrag V, Michot JM,
Zinzani PL, Kuruvilla J et al (2016) Programmed
death-1 blockade with pembrolizumab in patients
with ­
classical Hodgkin lymphoma after brentuximab vedotin failure. J Clin Oncol 34(31):
3733–3739
104. Pardoll DM (2012) The blockade of immune checkpoints in cancer immunotherapy. Nat Rev Cancer
12(4):252–264
105. Cheson BD, Ansell S, Schwartz L, Gordon LI,
Advani R, Jacene HA et al (2016) Refinement of
the Lugano classification lymphoma response crite-
A. Gallamini et al.
140
106.
107.
108.
109.
110.
111.
112.
113.
114.
115.
116.
ria in the era of immunomodulatory therapy. Blood
128(21):2489
Girinsky T, Pichenot C, Beaudre A, Ghalibafian
M, Lefkopoulos D (2006) Is intensity-modulated
radiotherapy better than conventional radiation
treatment and three-dimensional conformal radiotherapy for mediastinal masses in patients with
Hodgkin’s disease, and is there a role for beam orientation optimization and dose constraints assigned
to virtual volumes? Int J Radiat Oncol Biol Phys
64(1):218–226
Girinsky T, van der Maazen R, Specht L, Aleman B,
Poortmans P, Lievens Y et al (2006) Involved-node
radiotherapy (INRT) in patients with early Hodgkin
lymphoma: concepts and guidelines. Radiother
Oncol 79(3):270–277
Specht L, Gray RG, Clarke MJ, Peto R (1998)
Influence of more extensive radiotherapy and
adjuvant chemotherapy on long-term outcome of
early-stage Hodgkin's disease: a meta-analysis
of 23 randomized trials involving 3,888 patients.
International Hodgkin's disease collaborative group.
J Clin Oncol 16(3):830–843
Yahalom J (2005) Transformation in the use of radiation therapy of Hodgkin lymphoma: new concepts
and indications lead to modern field design and are
assisted by PET imaging and intensity modulated
radiation therapy (IMRT). Eur J Haematol Suppl
66:90–97
Gregoire V (2004) Is there any future in radiotherapy planning without the use of PET: unraveling the
myth. Radiother Oncol 73(3):261–263
Berthelsen AK, Dobbs J, Kjellén E, Landberg T,
Möller T, Nilsson P et al (2007) What’s new in
target volume definitions for radiologists in ICRU
report 71? How can the ICRU volume definitions
be integrated in clinical practice? Cancer Imaging
7(1):104–116
Jarritt PH, Carson KJ, Hounsell AR, Visvikis D
(2006) The role of PET/CT scanning in radiotherapy
planning. Br J Radiol 79(1):S27–S35
Specht L, Yahalom J, Illidge T, Berthelsen AK,
Constine LS, Eich HT et al (2014) Modern radiation therapy for Hodgkin lymphoma: field and dose
guidelines from the international lymphoma radiation oncology group (ILROG). Int J Radiat Oncol
Biol Phys 89(4):854–862
Specht L (2007) 2-[18F]fluoro-2-deoxyglucose
positron-emission tomography in staging, response
evaluation, and treatment planning of lymphomas.
Semin Radiat Oncol 17(3):190–197
van Baardwijk A, Baumert BG, Bosmans G, van
Kroonenburgh M, Stroobants S, Gregoire V et al
(2006) The current status of FDG-PET in tumour
volume definition in radiotherapy treatment planning. Cancer Treat Rev 32(4):245–260
Dizendorf EV, Baumert BG, von Schulthess
GK, Lutolf UM, Steinert HC (2003) Impact of
whole-body 18F-FDG PET on staging and managing patients for radiation therapy. J Nucl Med
44(1):24–29
117. Lee YK, Cook G, Flower MA, Rowbottom C,
Shahidi M, Sharma B et al (2004) Addition of
18F-FDG-PET scans to radiotherapy planning
of thoracic lymphoma. Radiother Oncol 73(3):
277–283
118. Hutchings M, Loft A, Hansen M, Berthelsen AK,
Specht L (2007) Clinical impact of FDG-PET/CT in
the planning of radiotherapy for early-stage Hodgkin
lymphoma. Eur J Haematol 78(3):206–212
119. Girinsky T, Ghalibafian M, Bonniaud G, Bayla A,
Magne N, Ferreira I et al (2007) Is FDG-PET scan in
patients with early stage Hodgkin lymphoma of any
value in the implementation of the involved-node
radiotherapy concept and dose painting? Radiother
Oncol 85(2):178–186
120. Josting A, Muller H, Borchmann P, Baars JW,
Metzner B, Dohner H et al (2010) Dose intensity of
chemotherapy in patients with relapsed Hodgkin’s
lymphoma. J Clin Oncol 28(34):5074–5080
121. Brockelmann PJ, Muller H, Casasnovas O, Hutchings
M, von TB, Jurgens M et al (2017) Risk factors and
a prognostic score for survival after autologous
stem-cell transplantation for relapsed or refractory
Hodgkin lymphoma. Ann Oncol 28(6):1352–1358
122. Moskowitz CH, Matasar MJ, Zelenetz AD,
Nimer SD, Gerecitano J, Hamlin P et al (2012)
Normalization of pre-ASCT, FDG-PET imaging with second-line, non-cross-resistant, chemotherapy programs improves event-free survival in
patients with Hodgkin lymphoma. Blood 119(7):
1665–1670
123. Gentzler RD, Evens AM, Rademaker AW, Weitner
BB, Mittal BB, Dillehay GL et al (2014) F-18 FDG-­
PET predicts outcomes for patients receiving total
lymphoid irradiation and autologous blood stem-cell
transplantation for relapsed and refractory Hodgkin
lymphoma. Br J Haematol 165(6):793–800
124. Devillier R, Coso D, Castagna L, Brenot R, Anastasia
A, Chiti A et al (2012) positron emission tomography response at the time of autologous stem cell
transplantation predicts outcome of patients with
relapsed and/or refractory Hodgkin's lymphoma
responding to prior salvage therapy. Haematologica
97(7):1073–1079
125. Poulou LS, Thanos L, Ziakas PD (2010) Unifying
the predictive value of pretransplant FDG PET in
patients with lymphoma: a review and meta-analysis
of published trials. Eur J Nucl Med Mol Imaging
37(1):156–162
126. Moskowitz AJ, Schoder H, Gavane S, Thoren KL,
Fleisher M, Yahalom J et al (2017) Prognostic significance of baseline metabolic tumor volume in
relapsed and refractory Hodgkin lymphoma. Blood
130(20):2196–2203
127. Gallamini A (2017) Relapsed/refractory HL: FDG-­
PET is the trump card. Blood 130(20):2154–2155
128. Armitage JO, Loberiza FR (2006) Is there a place for
routine imaging for patients in complete remission from
aggressive lymphoma? Ann Oncol 17(6):883–884
129. Radford JA, Eardley A, Woodman C, Crowther
D (1997) Follow up policy after treatment for
7
Functional Imaging in Hodgkin Lymphoma
130.
131.
132.
133.
134.
135.
136.
137.
138.
139.
140.
Hodgkin’s disease: too many clinic visits and
routine tests? A review of hospital records. BMJ
314(7077):343–346
Gallamini A, Kostakoglu L (2012) Positron emission
tomography/computed tomography surveillance in
patients with lymphoma: a fox hunt? Haematologica
97(6):797–799
Petrausch U, Samaras P, Veit-Haibach P, Tschopp
A, Soyka JD, Knuth A et al (2010) Hodgkin’s lymphoma in remission after first-line therapy: which
patients need FDG-PET/CT for follow-up? Ann
Oncol 21(5):1053–1057
Zinzani PL, Stefoni V, Tani M, Fanti S, Musuraca
G, Castellucci P et al (2009) Role of [18F]fluorodeoxyglucose positron emission tomography
scan in the follow-up of lymphoma. J Clin Oncol
27(11):1781–1787
Jerusalem G, Beguin Y, Fassotte MF, Belhocine T,
Hustinx R, Rigo P et al (2003) Early detection of
relapse by whole-body positron emission tomography in the follow-up of patients with Hodgkin's disease. Ann Oncol 14(1):123–130
El Galaly TC, Mylam KJ, Brown P, Specht L,
Christiansen I, Munksgaard L et al (2012) Positron
emission tomography/computed tomography surveillance in patients with Hodgkin lymphoma in first
remission has a low positive predictive value and
high costs. Haematologica 97(6):931–936
Aleman BM, van den Belt-Dusebout AW, Klokman
WJ, Van’t Veer MB, Bartelink H, van Leeuwen
FE (2003) Long-term cause-specific mortality of
patients treated for Hodgkin's disease. J Clin Oncol
21(18):3431–3439
Raemaekers JM, Andre MP, Federico M, Girinsky T,
Oumedaly R, Brusamolino E et al (2014) Omitting
radiotherapy in early positron emission tomography-­
negative stage I/II Hodgkin lymphoma is associated with an increased risk of early relapse: clinical
results of the preplanned interim analysis of the randomized EORTC/LYSA/FIL H10 trial. J Clin Oncol
32(12):1188–1194
Fuchs M, Goergen H, Kobe C, Kuhnert G, Lohri A,
Greil R et al (2019) Positron Emission Tomographyguided treatment in early-stage favorable Hodgkin
lymphoma: final results of the international, randomized phase III HD16 trial by the German Hodgkin
Study Group. J Clin Oncol 37(31):2835–2845
Engert A, Diehl V, Franklin J, Lohri A, Dorken B,
Ludwig WD et al (2009) Escalated-dose BEACOPP
in the treatment of patients with advanced-stage
Hodgkin's lymphoma: 10 years of follow-up of the
GHSG HD9 study. J Clin Oncol 27(27):4548–4554
Trotman J (2017) Foss+Ñ a, Federico M, Stevens
L, Kirkwood a, Clifton-Hadley L, et al. Response-­
adjusted therapy for advanced Hodgkin lymphoma
(rathl) trial: longer follow up confirms efficacy
of de-escalation after a negative interim PET scan
(CRUK/07/033). Hematol Oncol 35(S2):65–67
Borchmann P, Goergen H, Kobe C, Lohri A, Greil R,
Eichenauer DA et al (2018) PET-guided treatment in
patients with advanced-stage Hodgkin’s lymphoma
141
141.
142.
143.
144.
145.
146.
147.
148.
149.
150.
(HD18): final results of an open-label, international,
randomised phase 3 trial by the German Hodgkin
study group. Lancet 390(10114):2790–2802
Borchmann P, Goergen H, Kobe C, Lohri A, Greil
R, Eichenauer DA et al (2017) Early interim PET in
patients with advanced-stage Hodgkin’s lymphoma
treated within the phase 3 GHSG HD18 study. Blood
130(Suppl 1):737
Casasnovas RO, Bouabdallah R, Brice P, Lazarovici
J, Ghesquieres H, Stamatoullas A et al (2019) PET-­
adapted treatment for newly diagnosed advanced
Hodgkin lymphoma (AHL2011): a randomised,
multicentre, non-inferiority, phase 3 study. Lancet
Oncol 20(2):202–215
Kobe C, Dietlein M, Franklin J, Markova J, Lohri A,
Amthauer H et al (2008) Positron emission tomography has a high negative predictive value for progression or early relapse for patients with residual disease
after first-line chemotherapy in advanced-stage
Hodgkin lymphoma. Blood 112(10):3989–3994
Mocikova H, Pytlik R, Markova J, Steinerova K,
Kral Z, Belada D et al (2011) Pre-transplant positron emission tomography in relapsed Hodgkin lymphoma patients. Leuk Lymphoma 52:1668
Smeltzer JP, Cashen AF, Zhang Q, Homb A,
Dehdashti F, Abboud CN et al (2011) Prognostic
significance of FDG-PET in relapsed or refractory
classical Hodgkin lymphoma treated with standard salvage chemotherapy and autologous stem
cell transplantation. Biol Blood Marrow Transplant
17(11):1646–1652
Moskowitz AJ, Yahalom J, Kewalramani T,
Maragulia JC, Vanak JM, Zelenetz AD et al (2010)
Pretransplantation functional imaging predicts outcome following autologous stem cell transplantation for relapsed and refractory Hodgkin lymphoma.
Blood 116(23):4934–4937
Moskowitz AJ, Schoder H, Yahalom J, McCall SJ,
Fox SY, Gerecitano J et al (2015) PET-adapted
sequential salvage therapy with brentuximab vedotin followed by augmented ifosfamide, carboplatin,
and etoposide for patients with relapsed and refractory Hodgkin's lymphoma: a non-randomised, open-­
label, single-centre, phase 2 study. Lancet Oncol
16(3):284–292
Reyal Y, Kayani I, Bloor AJC, Fox CP, Chakraverty
R, Sjursen AM et al (2016) Impact of pretransplantation 18F-fluorodeoxyglucose-positron emission tomography on survival outcomes after T cell
depleted allogeneic transplantation for Hodgkin
lymphoma. Biol Blood Marrow Transplant
22(7):1234–1241
Marani C, Raiola AM, Morbelli S, Dominietto A,
Ferrarazzo G, Avenoso D et al (2018) Haploientical
transplants with post-transplant cyclophosphamide
for relapsed or refractory Hodgkin lymphoma: the
role of comorbidity index and Pretransplant positron emission tomography. Biol Blood Marrow
Transplant 24(12):2501–2508
Dodero A, Crocchiolo R, Patriarca F, Miceli R,
Castagna L, Ciceri F et al (2010) Pretransplantation
A. Gallamini et al.
142
151.
152.
153.
154.
155.
156.
157.
158.
159.
160.
[18-F]fluorodeoxyglucose positron emission tomography scan predicts outcome in patients with recurrent Hodgkin lymphoma or aggressive non-Hodgkin
lymphoma undergoing reduced-intensity conditioning followed by allogeneic stem cell transplantation.
Cancer 116(21):5001–5011
Lambert JR, Bomanji JB, Peggs KS, Thomson
KJ, Chakraverty RK, Fielding AK et al (2010)
Prognostic role of PET scanning before and after
reduced-intensity allogeneic stem cell transplantation for lymphoma. Blood 115(14):2763–2768
Hart DP, Avivi I, Thomson KJ, Peggs KS, Morris
EC, Goldstone AH et al (2005) Use of 18F-FDG
positron emission tomography following allogeneic
transplantation to guide adoptive immunotherapy
with donor lymphocyte infusions. Br J Haematol
128(6):824–829
Juweid ME, Wiseman GA, Vose JM, Ritchie JM,
Menda Y, Wooldridge JE et al (2005) Response
assessment of aggressive non-Hodgkin's lymphoma
by integrated international workshop criteria and
fluorine-18-fluorodeoxyglucose positron emission
tomography. J Clin Oncol 23(21):4652–4661
Cheson BD, Horning SJ, Coiffier B, Shipp MA,
Fisher RI, Connors JM et al (1999) Report of an
international workshop to standardize response criteria for non-Hodgkin's lymphomas. NCI Sponsored
International Working Group. J Clin Oncol
17(4):1244
Trotman J, Luminari S, Boussetta S, Versari A,
Dupuis J, Tychyj C et al (2014) Prognostic value of
PET-CT after first-line therapy in patients with follicular lymphoma: a pooled analysis of central scan
review in three multicentre studies. Lancet Haematol
1(1):e17–e27
Puccini B, Nassi L, Minoia C, Volpetti S, Ciancia
R, Riccomagno PC et al (2017) Role of bone marrow biopsy in staging of patients with classical
Hodgkin's lymphoma undergoing positron emission
tomography/computed tomography. Ann Hematol
96(7):1147–1153
Eve HE, Rule SA (2010) Lenalidomide-induced
tumour flare reaction in mantle cell lymphoma. Br J
Haematol 151(4):410–412
Corazzelli G, De FR, Capobianco G, Frigeri F, De
R et al (2010) Tumor flare reactions and response
to lenalidomide in patients with refractory classic Hodgkin lymphoma. Am J Hematol 85(1):
87–90
Andritsos LA, Johnson AJ, Lozanski G, Blum W,
Kefauver C, Awan F et al (2008) Higher doses of
lenalidomide are associated with unacceptable toxicity including life-threatening tumor flare in patients
with chronic lymphocytic leukemia. J Clin Oncol
26(15):2519–2525
Chanan-Khan A, Miller KC, Musial L, Lawrence D,
Padmanabhan S, Takeshita K et al (2006) Clinical
efficacy of lenalidomide in patients with relapsed or
refractory chronic lymphocytic leukemia: results of
a phase II study. J Clin Oncol 24(34):5343–5349
161. Han HS, Escalon MP, Hsiao B, Serafini A, Lossos IS
(2009) High incidence of false-positive PET scans in
patients with aggressive non-Hodgkin's lymphoma
treated with rituximab-containing regimens. Ann
Oncol 20(2):309–318
162. Skoura E, Ardeshna K, Halsey R, Wan S, Kayani I
(2016) False-positive 18F-FDG PET/CT imaging:
dramatic “flare response” after rituximab administration. Clin Nucl Med 41(3):e171–e172
163. Pro B, Advani R, Brice P, Bartlett NL, Rosenblatt
JD, Illidge T et al (2012) Brentuximab vedotin
(SGN-35) in patients with relapsed or refractory systemic anaplastic large-cell lymphoma: results of a
phase II study. J Clin Oncol 30(18):2190–2196
164. Gopal AK, Schuster SJ, Fowler NH, Trotman J,
Hess G, Hou JZ et al (2018) Ibrutinib as treatment for patients with relapsed/refractory follicular lymphoma: results from the open-label,
multicenter, phase II DAWN study. J Clin Oncol
36(23):2405–2412
165. Brown JR, Byrd JC, Coutre SE, Benson DM, Flinn
IW, Wagner-Johnston ND et al (2014) Idelalisib, an
inhibitor of phosphatidylinositol 3-kinase p110delta,
for relapsed/refractory chronic lymphocytic leukemia. Blood 123(22):3390–3397
166. Kurtz DM, Green MR, Bratman SV, Scherer F, Liu
CL, Kunder CA et al (2015) Noninvasive monitoring of diffuse large B-cell lymphoma by immunoglobulin high-throughput sequencing. Blood
125(24):3679–3687
167. van Eijndhoven MA, Zijlstra JM, Groenewegen NJ,
Drees EE et al (2016) Plasma vesicle miRNAs for
therapy response monitoring in Hodgkin lymphoma
patients. JCI Insight 1(19):e89631
168. Spina V, Bruscaggin A, Cuccaro A, Martini M, Di
TM, Forestieri G et al (2018) Circulating tumor
DNA reveals genetics, clonal evolution, and residual disease in classical Hodgkin lymphoma. Blood
131(22):2413–2425
169. Cuccaro A, Annunziata S, Cupelli E, Martini M,
Calcagni ML, Rufini V et al (2016) CD68+ cell
count, early evaluation with PET and plasma TARC
levels predict response in Hodgkin lymphoma.
Cancer Med 5(3):398–406
170. Agostinelli C, Gallamini A, Stracqualursi L, Agati
P, Tripodo C, Fuligni F et al (2016) The combined
role of biomarkers and interim PET scan in prediction of treatment outcome in classical Hodgkin's
lymphoma: a retrospective, European, Multicentre
Cohort Study. Lancet Haematol 3(10):e467–e479
171. Lin C, Itti E, Haioun C, Petegnief Y, Luciani A,
Dupuis J et al (2007) Early 18F-FDG PET for prediction of prognosis in patients with diffuse large
B-cell lymphoma: SUV-based assessment versus
visual analysis. J Nucl Med 48(10):1626–1632
172. Itti E, Lin C, Dupuis J, Paone G, Capacchione D,
Rahmouni A et al (2009) Prognostic value of interim
18F-FDG PET in patients with diffuse large B-cell
lymphoma: SUV-based assessment at 4 cycles of
chemotherapy. J Nucl Med 50(4):527–533
7
Functional Imaging in Hodgkin Lymphoma
173. Biggi A, Bergesio F, Chauvie S, Bianchi A, Menga
M, Fallanca F et al (2017) Concomitant semi-­
quantitative and visual analysis improves the predictive value on treatment outcome of interim
18F-fluorodeoxyglucose/positron emission tomography in advanced Hodgkin lymphoma. Q J Nucl
Med Mol Imaging. https://doi.org/10.23736/
S1824-4785
174. Cerci JJ, Trindade E, Pracchia LF, Pitella FA,
Linardi CC, Soares J Jr et al (2010) Cost effectiveness of positron emission tomography in patients
with Hodgkin’s lymphoma in unconfirmed complete
remission or partial remission after first-line therapy.
J Clin Oncol 28(8):1415–1421
175. Kobe C, Kuhnert G, Kahraman D, Haverkamp H,
Eich HT, Franke M et al (2014) Assessment of tumor
size reduction improves outcome prediction of positron emission tomography/computed tomography
after chemotherapy in advanced-stage Hodgkin lymphoma. J Clin Oncol 32(17):1776–1781
176. Al-Nabhani KZ, Syed RF, Michopoulou S, Alkalbani
J, Alkalbani JF, Afaq AF, Panagiotidis E, Meara C
et al (2014) Qualitative and quantitative comparison
of PET/CT and PET/MR imaging in clinical practice. J Nucl Med 55(1):88–94
177. Stecco AA, Buemi F, Iannessi AA, Carriero AA,
Gallamini A (2018) Current concepts in tumor imaging with whole-body MRI with diffusion imaging
(WB-MRI-DWI) in multiple myeloma and lymphoma. Leuk Lymphoma 59(11):2546–2556
178. Herrmann K, Queiroz M, Huellner MW, de Galiza
BF, Buck A, Schaefer N et al (2015) Diagnostic performance of FDG-PET/MRI and WB-DW-MRI in
the evaluation of lymphoma: a prospective comparison to standard FDG-PET/CT. BMC Cancer 15:1002
179. Martiat P, Ferrant A, Labar D, Cogneau M, Bol
A, Michel C et al (1988) In vivo measurement of
carbon-­11 thymidine uptake in non-Hodgkin’s lymphoma using positron emission tomography. J Nucl
Med 29(10):1633–1637
180. Shields AF, Mankoff DA, Link JM, Graham MM,
Eary JF, Kozawa SM et al (1998) Carbon-11-­
thymidine and FDG to measure therapy response. J
Nucl Med 39(10):1757–1762
181. Shields AF, Grierson JR, Dohmen BM, Machulla
HJ, Stayanoff JC, Lawhorn-Crews JM et al
(1998) Imaging proliferation in vivo with [F-18]
FLT and positron emission tomography. Nat Med
4(11):1334–1336
182. Buchmann I, Neumaier B, Schreckenberger M,
Reske S (2004) [18F]3′-deoxy-3′-fluorothymidine-­
PET in NHL patients: whole-body biodistribution
and imaging of lymphoma manifestations—a pilot
study. Cancer Biother Radiopharm 19(4):436–442
183. Buck AK, Bommer M, Stilgenbauer S, Juweid M,
Glatting G, Schirrmeister H et al (2006) Molecular
imaging of proliferation in malignant lymphoma.
Cancer Res 66(22):11055–11061
184. Kasper B, Egerer G, Gronkowski M, Haufe S,
Lehnert T, Eisenhut M et al (2007) Functional
143
185.
186.
187.
188.
189.
190.
191.
192.
193.
194.
195.
diagnosis of residual lymphomas after radiochemotherapy with positron emission tomography
comparing FDG- and FLT-PET. Leuk Lymphoma
48(4):746–753
Graf N, Herrmann K, den Hollander J, Fend F,
Schuster T, Wester HJ et al (2008) Imaging proliferation to monitor early response of lymphoma to cytotoxic treatment. Mol Imaging Biol 10(6):349–355
Stern PH, Wallace CD, Hoffman RM (1984) Altered
methionine metabolism occurs in all members of a
set of diverse human tumor cell lines. J Cell Physiol
119(1):29–34
Nuutinen J, Leskinen S, Lindholm P, Soderstrom
KO, Nagren K, Huhtala S et al (1998) Use of carbon-11 methionine positron emission tomography
to assess malignancy grade and predict survival in
patients with lymphomas. Eur J Nucl Med 25(7):
729–735
Kaste SC, Howard SC, McCarville EB, Krasin MJ,
Kogos PG, Hudson MM (2005) 18F-FDG-avid sites
mimicking active disease in pediatric Hodgkin’s.
Pediatr Radiol 35(2):141–154
Kawase Y, Yamamoto Y, Kameyama R, Kawai N,
Kudomi N, Nishiyama Y (2011) Comparison of
11C-methionine PET and 18F-FDG PET in patients
with primary central nervous system lymphoma.
Mol Imaging Biol 13(6):1284–1289
Rylova SN, Del PL, Klingeberg C, Tonnesmann
R, Illert AL, Meyer PT et al (2016) Immuno-­
PET imaging of CD30-positive lymphoma using
89Zr-desferrioxamine-labeled CD30-specific AC-10
antibody. J Nucl Med 57(1):96–102
Rossi C, Kanoun S, Berriolo-Riedinger A, Dygai-­
Cochet I, Humbert O, Legouge C et al (2014)
Interim 18F-FDG PET SUVmax reduction is
superior to visual analysis in predicting outcome
early in Hodgkin lymphoma patients. J Nucl Med
55(4):569–573
Specht L, Nordentoft AM, Cold S, Clausen NT,
Nissen NI (1988) Tumor burden as the most important prognostic factor in early stage Hodgkin’s
disease. Relations to other prognostic factors
and implications for choice of treatment. Cancer
61(8):1719–1727
Song MK, Chung JS, Lee JJ, Jeong SY, Lee SM,
Hong JS et al (2013) Metabolic tumor volume by
positron emission tomography/computed tomography as a clinical parameter to determine therapeutic modality for early stage Hodgkin’s lymphoma.
Cancer Sci 104(12):1656–1661
Kanoun S, Rossi C, Berriolo-Riedinger A, Dygai-­
Cochet I, Cochet A, Humbert O et al (2014) Baseline
metabolic tumour volume is an independent prognostic factor in Hodgkin lymphoma. Eur J Nucl Med
Mol Imaging 41(9):1735–1743
Mettler J, Muller H, Voltin CA, Baues C, Klaeser B,
Moccia A et al (2018) Metabolic tumour volume for
response prediction in advanced-stage Hodgkin lymphoma. J Nucl Med 2018:pii: jnumed.118.210047
8
Prognostic Factors
Paul J. Bröckelmann and Lena Specht
Contents
8.1
Historical Perspective 146
8.2
8.2.1
8.2.2
8.2.3
8.2.4
8.2.5
Prognostic Factors Definition and Use Types of Prognostic Factors Different End Points Types and Analyses of Prognostic Studies Predictive Factors 147
147
147
148
148
148
8.3
Prognostic Factors across Stages 149
8.4
Early-Stage Hodgkin Lymphoma 149
8.5
8.5.1
8.5.2
Advanced-Stage Hodgkin Lymphoma retreatment Prognostic Factors in Advanced-Stage Hodgkin Lymphoma P
Interim PET as Prognostic Factor in Advanced-Stage Hodgkin
Lymphoma Prognostic Indices in Advanced-Stage Hodgkin Lymphoma 150
152
8.5.3
8.6
8.6.1
8.6.2
8.6.3
rognostic Factors in Relapsed or Refractory
P
Hodgkin Lymphoma Patients Treated for r/r HL with Conventional Treatment Treatment with High-Dose Chemotherapy and Autologous
Stem Cell Transplantation Patients’ Prognostic Indices or Scores in r/r HL
Treated with Novel Agents 153
153
155
155
157
158
P. J. Bröckelmann (*)
Department I of Internal Medicine and German
Hodgkin Study Group (GHSG),
University Hospital of Cologne, Köln, Germany
e-mail: paul.broeckelmann@uk-koeln.de
L. Specht
Department of Oncology, Section 3994, The Finsen
Centre, Rigshospitalet, University of Copenhagen,
Copenhagen, Denmark
e-mail: lena.specht@regionh.dk
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_8
145
P. J. Bröckelmann and L. Specht
146
8.6.4
Prognostic Factors for Treatment with Novel Agents in r/r HL 158
8.7
Conclusion and Future Aspects 159
References Abbreviations
ABVD
Adriamycin, bleomycin, vinblastine, dacarbazine
ASCT
Autologous stem cell transplant
BEACOPPescBleomycin, etoposide, adriamycin, cyclophosphamide, vincristine,
procarbazine,
prednisolone, escalated
BNLI
British National Lymphoma
Investigation
BV
Brentuximab vedotin
CALGB
Cancer and Leukemia Group B
cfDNA
Cell-free DNA
CRP
C-reactive protein
CS
Clinical stage
EBMTEuropean Society for Blood and
Marrow Transplantation
ECOGEastern Cooperative Oncology
Group
EFS
Event-free survival
EN
Extranodal
EORTC
European Organization for
Research and Treatment of
Cancer
ESR
Erythrocyte sedimentation rate
FDG2-[18F]fluoro-2-deoxy-d-glucose
FF2F
Freedom from second failure
FFP
Freedom from progression
GELA
Groupe
d’Etudes
des
Lymphomes de l’Adulte
GHSG
German Hodgkin Lymphoma
Study Group
HDCT
High-dose chemotherapy
HL
Hodgkin lymphoma
HRS
Hodgkin-Reed-Sternberg
IPS
International Prognostic Score
LDH
Lactate dehydrogenase
LMM
Large mediastinal mass
LP
Lymphocyte predominant
160
MHCMajor
histocompatibility
complex
MOPPMechlorethamine, vincristine,
procarbazine, prednisolone
MTV
Metabolic tumour volume
NCI-C
National Cancer Institute of
Canada
NCI-USNational Cancer Institute of the
United States
NS
Nodular sclerosis
OS
Overall survival
PD-1
Programmed cell death protein 1
PD-L1
Programmed death ligand 1
PET
Positron emission tomography
PFS
Progression-free survival
PR
Partial remission
r/r
Relapsed or refractory
RT
Radiation therapy
SWOG
Southwest Oncology Group
TAM
Tumour-associated macrophages
TARCThymus and activation regulated
chemokine
TNF
Tumour necrosis factor
TTR
Time to relapse
8.1
Historical Perspective
The concept that Hodgkin lymphoma (HL, then
called Hodgkin’s disease) passes through successive clinical stages with increasing spread of the
disease and progressive worsening of prognosis
was developed early on [1]. Different staging
classifications were proposed based on the anatomic extent of disease [2–8]. A consensus was
reached at the Workshop on the Staging of
Hodgkin’s Disease at Ann Arbor in 1971 [9], and
the Ann Arbor staging classification was universally adopted. It still remains the basis for the
Prognostic Factors
Fig. 8.1 Five-year
relative survival
according to Ann Arbor
stage for patients in the
US National Cancer
Institute’s Surveillance,
Epidemiology, and End
Results (SEER) program
treated in the period
1975–2015 (Reprinted
with permission from
Noone et al. [18])
147
5-Year Relative Survival
100
92.3%
93.4%
90
83.0%
80
Percent Surviving
8
82.7%
72.9%
70
60
50
40
30
20
10
0
Stage I
Stage II
Stage III
Stage IV
Unknown
Stage
evaluation of patients with HL, and its prognostic
significance has been documented in numerous
studies of patients treated with different treatment modalities [10–17]. Five-year relative survival according to Ann Arbor stage for patients
treated between 1975 and 2015 from the US
National Cancer Institute’s Surveillance,
Epidemiology, and End Results (SEER) program
is shown in Fig. 8.1 [18].
However, the extent of disease varies within the
Ann Arbor stages leading to variations in prognosis. A modification of the Ann Arbor classification
was proposed at the Cotswold meeting, incorporating a designation for number of sites and bulk [19].
This modification has not been universally adopted.
An update of the staging system was introduced in
2014 with the Lugano classification [20]. Therein a
single nodal mass of 10 cm or greater than a third
of the transthoracic diameter at any level of thoracic vertebrae as determined by CT was retained
as the definition of bulky disease. PET/CT was designated as the preferred method of staging, obviating the need for bone marrow biopsies.
Numerous other prognostic factors for different Ann Arbor stages, disease presentations,
treatment approaches, and outcomes have been
introduced, and varying combinations of these
factors are currently being used by different
centres and study groups.
8.2
Prognostic Factors
8.2.1
Definition and Use
Prognostic factors are variables measured in
individual patients that offer a partial explanation of the outcome heterogeneity of a given disease [21]. They are important in clinical practice
for allocating patients into different risk groups,
for selection of treatment strategy, and as an aid
in patient counselling [22]. However, it is important to realise that prediction is very uncertain
for the individual patient. Statements of probability can be made, but even these will be more
accurate for groups of patients than for individuals [23]. Prognostic factors can also be used in
the design of clinical trials to define eligibility
criteria and strata to ensure comparability of
treatment groups [21–24]. However, prognostic
factors are rarely sufficiently explanatory to justify the comparison of treatments by use of nonrandomised data [25, 26].
8.2.2
Types of Prognostic Factors
Prognostic factors are divided into tumour-­related
factors, host-related factors, and environment-­
related factors [22]. Tumour-related factors
P. J. Bröckelmann and L. Specht
148
include those directly related to the presence of
the tumour or its effect on the host, reflecting
tumour pathology, anatomic extent, or tumour
biology. Host-related factors include factors that
are not directly related to the tumour but which
may significantly influence outcome, such as
demographic characteristics and co-morbidity.
Environment-related factors include factors
­outside the patient, such as socio-economic status
and access to and quality of health care.
The values of prognostic factors are generally
assumed to be known from the outset, before start
of treatment, so-called fixed covariates. However,
other important prognostic variables may only be
known later, such as time to response, toxicity of
treatment, and the value of presumed markers.
These are time-dependent covariates. They may
be important for answering biological questions,
but they should not be applied for adjustment for
treatment comparison, as they are themselves
affected by treatment [21–23].
8.2.3
Different End Points
Different outcomes may be of interest in analyses
of prognostic factors. Overall survival and
progression-­free survival are usually analysed,
but others may be relevant, e.g. disease-free survival for early-stage patients as nearly all patients
achieve remission. For each end point, there must
be clear information on the point in time from
which it is measured and the clinical characteristics of events and censoring. International guidelines have been published [27].
8.2.4
Types and Analyses
of Prognostic Studies
Three different study phases of prognostic factors
have been proposed, beginning with phase I early
exploratory analyses to identify potential markers and generate hypotheses for further investigation. Phase II studies are exploratory studies
attempting to use values of a proposed prognostic
factor to discriminate between high- and low-risk
patients. Phase III studies are large, confirmatory
studies based on prespecified hypotheses involving one or a few new factors, and the purpose of
these studies is to determine how much the new
factor adds to the predictive power of already
accepted factors [24, 28].
A useful prognostic factor must be significant, independent, and clinically important [29].
Many variables may be prognostic in univariate
analysis. However, different variables are likely
to be interrelated. The important question is
whether a particular variable adds useful information to what is already known. Multiple
regression analysis is commonly employed to
determine whether a variable has independent
significance when other known variables are
taken into account. This kind of analysis may
form the basis for the development of a prognostic model and a risk score or risk groups
[28]. The Cox proportional hazards regression
model is most commonly used when time-toevent outcomes are of interest [30]. The selection of variables for the final model is usually
done by stepwise selection. By play of chance,
different factors may be selected in different
studies. An important additional analysis for a
new marker is therefore to determine its prognostic ability in a model including all previously
defined prognostic factors [28, 31]. Differences
may also be due to small sample size, different
assay techniques, different cut points for variables, inclusion of different subsets of patients,
and different study end points.
8.2.5
Predictive Factors
Predictive factors are patient characteristics that
identify subgroups of patients with different outcomes as a consequence of a given treatment.
Hence a predictive factor identifies subgroups of
treated patients having different outcomes,
whereas a prognostic factor identifies subgroups
of untreated patients having different outcomes.
A factor that is predictive of outcome after one
particular treatment may not be predictive for
another treatment.
8
Prognostic Factors
8.3
149
rognostic Factors across
P
Stages
[52–55]. Hopefully this problem will be solved
soon, resulting in a simple method that will allow
measuring the total metabolic tumour volume.
This should be closely related to the total tumour
burden and become a practical and clinically useful method for dividing patients with HL into
prognostic subgroups (Fig. 8.2).
Most important prognostic factors in HL, irrespective of stage, are correlated with and provide
indirect measures of the patient’s total tumour
burden. The total tumour burden is the most
important prognostic factor, rendering most other
prognostic factors insignificant in multivariate
analysis. This was demonstrated by semi-­ 8.4
Early-Stage Hodgkin
quantitative methods, based on a combination of
Lymphoma
the number of involved lymph node regions and
the volume of disease in individual regions, in From early studies in Hodgkin lymphoma, it was
patient materials treated even before staging with evident that the number of involved regions and
CT scans was introduced [32–39]. When patient size of mediastinal disease, B symptoms, histomaterials staged with CT scans became available, logical subtype, age, gender, ESR, haemoglobin,
it became possible to directly contour and quanti- and serum albumin were prognostically signifitate the total tumour volume in each individual cant [14, 56–65]. Table 8.1 lists the established
patient, and the pivotal prognostic role of the
total tumour burden was confirmed [40–46].
However, these methods for estimating the total Table 8.1 Prognostic factors in early-stage HL
tumour burden were too laborious to be imple- Tumour burden
Number of involved lymph node regions
mented into clinical practice.
HL is a highly FDG-avid tumour, and the total Large tumour mass, particularly mediastinal
B symptoms
volume of the PET-positive lymphoma tissue (the
Histological subtype
metabolic tumour volume) is highly correlated Age
with the total tumour burden. Hence, studies are Gender
now demonstrating the prognostic relevance of Erythrocyte sedimentation rate (ESR)
the metabolic tumour volume [47–51]. The pre- Haemoglobin
ferred method for measuring the baseline meta- Serum albumin
bolic tumour volume has not yet been agreed (early interim FDG-PET scan)
100
TMTV0 ≤225ml
80
DSS probability (%)
PFS probability (%)
100
60
40
TMTV0 >225ml
20
TMTV0 ≤225ml
80
60
40
TMTV0 >225ml
20
P= 0.001
P= 0.0015
0
0
0
10
20
30
40
50
Time (months)
60
70
Fig. 8.2 Progression-free (PFS) and disease-specific survival (DSS) for 59 patients with all stages of HL divided
into low and high total metabolic tumour volume, i.e. < or
0
10
20
30
40
50
Time (months)
60
70
>225 ml (Reprinted with permission from Kanoun et al.
2014 [49])
P. J. Bröckelmann and L. Specht
150
prognostic factors in early-stage HL. Based on
the most important factors, different centres and
groups divided early-stage patients into favourable and unfavourable. Table 8.2 shows the factors and risk groups defined by some of the major
groups. Although the factors included in these
risk group definitions are fairly similar, the risk
group definitions vary making comparisons
between patient materials difficult.
Treatment of early-stage patients is tailored
according to prognostic subgroups. Thus, in
many studies patients are selected, making the
detection of prognostic factors difficult. Most of
the important prognostic factors are correlated
providing indirect measures of the patient’s total
tumour burden [36, 39, 62].
Functional imaging with FDG-PET has become
an important part of staging and treatment evaluation of lymphomas. An early interim FDG-PET
scan after one or two cycles of chemotherapy has
been shown to be highly predictive of outcome
after combined modality treatment [66–71].
However, in early-stage disease, there are many
false-positive results, the predictive value depends
on the chemotherapy regimen used, and the majority of the patients with interim PET positivity were
cured with combined modality therapy, yielding a
positive predictive value of only 15% [72]. The
negative predictive value is very high in earlystage disease, which would be expected in a disease with a very good prognosis. In fact, because
over 90% of patients with early-­stage disease are
cured with standard treatment, the added information from a negative interim PET scan is actually
very limited [73, 74]. The early interim FDG-PET
scan may be regarded as an in vivo test of the chemosensitivity of disease. As the result of the scan
is not known at the outset, there is a methodological problem with this test. Strictly speaking, outcome according to the result of an early interim
FDG-PET scan should only be measured from the
time when it is available, and it should be regarded
more as a predictive factor indicating the sensitivity to a particular treatment rather than as a usual
prognostic factor.
From a clinical point of view, it would be better to predict the outcome with a given regimen
up front rather than having to initially administer
possibly ineffective or too intensive treatment.
The total metabolic tumour volume (MTV) as a
surrogate for the total tumour burden seems
promising in this respect. Furthermore, recent
research into molecular abnormalities in either
tumour cells or non-malignant background cells
has demonstrated numerous biomarkers, some
of which will hopefully allow individualization
of treatment up front [75–91]. However, for a
new prognostic factor to be proven useful, it
needs to be demonstrated to show independent
significance when the already known factors, in
particular the total tumour burden, are taken into
account.
The standard treatment for early-stage disease
is combined modality treatment. Recently, chemotherapy alone has been used, resulting in a
moderate increase in the risk of relapse.
Prognostic factors in this group of patients have
not been analysed in detail as large cohorts of
patients with reasonable follow-up are not yet
available.
8.5
Advanced-Stage Hodgkin
Lymphoma
Patients presenting in stage III/IV are universally
considered as advanced-stage requiring systemic
treatment, eventually followed by consolidative
RT in selected cases. Patients with stage IIB and
additional risk factors such as extranodal disease
(EN) or a large mediastinal mass (LMM) are
additionally regarded advanced-stage by some
study groups or centres for the purpose of first-­
line therapy.
Large data sets are important to reliably assess
the independent contributions of single routinely
documented clinical prognostic factors and two
very large data sets resulted from international
cooperation: The International Database on
Hodgkin’s Disease set up in 1989 combined
>14,000 individual patient data across all stages
from 20 study groups in the MOPP era [14]. In
1995, the International Prognostic Factors Project
on advanced Hodgkin lymphoma combined data
of 5141 advanced-stage patients mainly treated
with doxorubicin-containing regimen [92]. More
CS I or CS IIA with ≥1 risk factors
CS IIB with (c) or (d) but without (a) and (b)
GHSG
(a) Large mediastinal mass
(b) Extranodal disease
(c) ESR ≥ 50 without B symptoms or ≥30 with B
symptoms
(d) ≥3 nodal areas
CS I–II without risk factors
EORTC
(a) Large mediastinal mass
(b) Age ≥ 50 years
(c) ESR ≥ 50 without B symptoms or ≥with B
symptoms
(d) ≥4 nodal areas
CS I–II (supradiaphragmatic) without risk
factors
CS I–II (supradiaphragmatic) with ≥1 risk
factors
NCI-C and ECOG
(a) Histology other than
LP/NS
(b) Age ≥ 40 years
(c) ESR ≥ 50
(d) ≥4 nodal areas
CS I–II without risk
factors
CS I–II with ≥1 risk
factors
CS I–II without risk
factors
CS I–II with ≥1 risk
factors
Stanford
(a) B symptoms
(b) L
arge mediastinal
mass
GHSG German Hodgkin Study Group, EORTC European Organization for Research and Treatment of Cancer, NCI-C National Cancer Institute of Canada, ECOG Eastern
Cooperative Oncology Group
Unfavourable
Favourable
Risk factors
Table 8.2 Criteria for favourable vs. unfavourable early-stage HL as defined by different centres and cooperative study groups
8
Prognostic Factors
151
P. J. Bröckelmann and L. Specht
152
recent studies focussed on disease activity measured by interim and/or end-of-treatment PET to
further individualize treatment.
8.5.1
Pretreatment Prognostic
Factors in Advanced-Stage
Hodgkin Lymphoma
The most important patient-related prognostic factor for overall survival (OS) in advanced-stage HL
is age [93–97]. Elderly patients (>60 years) are
often excluded from clinical trials [98]. Prevalence
of co-morbidity is associated with age and treatment-related mortality is increased [99]. In patients
up to 65 years, age >45 years is an independent
prognostic factor for freedom from progression.
The impact of age is amplified in overall survival
as compared to progression-free survival due to
compromised results of salvage treatment in
elderly relapsed patients [100]. Male gender is
another independent, although quantitatively moderate, adverse patient-related prognostic factor
within advanced stages [14, 92, 101, 102].
In terms of disease-related factors, tumour
burden either quantified from conventional or
PET imaging is a main determinant of prognosis
[39–41, 62, 103]. As previously outlined, MTV is
increasingly being recognised as pre-therapeutic
prognostic factor but still rarely measured routinely. Stage IV marks dissemination of the disease to extranodal sites and is independently
prognostic within advanced-stage disease [14,
92]. A particularly bad prognosis for bone marrow, lung, or liver involvement could not be confirmed. It thus remains controversial whether
certain disease locations or the number of
involved extranodal sites carries independent
prognostic value within stage IV disease.
Several haematological and biochemical laboratory parameters form a cluster of interrelated
prognostic indicators that mirror both tumour
burden and inflammatory processes [75]. These
factors include decreased serum albumin and
haemoglobin levels as well as an elevated ESR,
LDH, C-reactive protein (CRP), and alkaline
phosphatase, which are correlated with one other
as well as with the presence of B symptoms and
tumour burden. Leukocyte and lymphocyte
counts form a second correlation cluster of laboratory parameters; in addition, leucocytosis and
lymphocytopenia have an independent prognostic impact [92].
A plethora of biological parameters such as
levels of cytokines released by HRS cells, soluble
forms of membrane-derived antigens, and cell-­
free tumour DNA have been investigated for
prognostic value not confined to advanced-stage
disease. Many of these studies were done in
rather small data sets, and findings are often not
yet integrated with other clinical or imaging factors in multivariate analysis. The soluble form of
the CD30 molecule (sCD30) is released by HRS
cells and is detectable in the serum of virtually all
untreated patients. It maintains independent
prognostic significance in multivariate analysis
in moderately sized data sets [78, 104–106]. The
thymus and activation-regulated chemokine
(TARC or CCL117) plays a role in the pathogenesis of HL and correlates with disease burden and
outcome [107]. A correlative analysis of the
recent randomized phase III ECHELON-1 trial
comparing ABVD and BV-AVD could not confirm a prognostic role for sCD30 or TARC since
neither of the two parameters correlated with
early or end-of-treatment response nor PFS
[108]. The relevance of cytokine levels such as
IL-1RA, IL-2R, IL-6, IL-10, or tumour necrosis
factor (TNF) requires further investigation [78,
109, 110] as does the role of elevated β2-­
microglobulin [111]. Circulating cell-free tumour
DNA (cfDNA) recently gained attention across
various tumour types since it potentially allows
minimally invasive monitoring of disease activity
over the course of treatment as well as genetic
assessment of the malignancy [112]. After a
proof-of-concept study for HL was published
[113], very recent studies highlight the prognostic potential of cfDNA, especially in correlation
with disease status by PET [114, 115].
Focussing on tumour architecture and composition, several histopathological markers in
tumour tissue either expressed by HRS or reactive bystander cells were identified. High density
of CD68+ macrophages has been shown to be
adversely prognostic with ABVD treatment [89,
8
Prognostic Factors
153
116]. Tumour-associated macrophages (TAM)
together with cells involved in a TH1 immune
response possibly create an inflammatory
­environment favouring rapid lymphoma proliferation. A recent analysis revealed frequent 9p24.1
amplifications in advanced-stage cHL which
were associated with poorer PFS in ABVD-­
treated patients [117]. In order to obtain a method
usable in clinical practice, a 23-gene expression
signature was shown to be predictive in advanced-­
stage ABVD-treated patients [88]. However, further validation in clinical trials including
PET-adapted or BEACOPP-based therapy is
required.
8.5.2
I nterim PET as Prognostic
Factor in Advanced-Stage
Hodgkin Lymphoma
An early interim FDG-PET scan after two cycles
of chemotherapy (PET-2) has been shown to be
highly predictive of outcome in advanced-stage
HL a decade ago [68, 118–120]. In a large study
of advanced-stage patients treated with ABVD,
the prognostic value of PET-2 completely overshadowed the role of the International Prognostic
Score (IPS) [121]. Figure 8.3 shows PFS according to the IPS and the PET-2 result.
Within the last years, individualized treatment
guided by PET-2 was investigated in several
phase III trials. Within the UK-led RATHL trial,
8.5.3
Progression-Free Survival
1.0
0.8
0.6
IPS 0-2, PET2 negative
IPS 0-2, PET2 positive
IPS 3-7, PET2 negative
IPS 3-7, PET2 positive
0.4
0.2
Log-rank P = ≈ 0
0
PFS and OS were superior in PET-2-negative
patients as compared to PET-2-positive patients,
despite de-escalation from ABVD to AVD or
escalation from ABVD to BEACOPPescalated,
respectively [122]. In PET-2-negative patients,
only stage IV disease and the IPS but no other
factors such as bulky disease, PET score by
Deauville, or B symptoms retained prognostic
significance for PFS. In contrast, PET-2 status
was not prognostic for PFS in the GHSG HD18
trial [123] as shown in Fig. 8.4.
Neither
addition
of
rituximab
to
BEACOPPescalated in the PET-2-positive patient
nor reduction from 6–8×BEACOPPescalated to
4×BEACOPPescalated did relevantly change PFS or
OS for these different treatment arms (Fig. 8.5)
[123]. Similar observations were made in the
French AHL2011 trial, which investigated de-­
escalation to 4xABVD in PET-2-negative
patients after 2×BEACOPPescalated [124]. These
findings underline the poor positive predictive
value of a positive PET-2 in patients initially
treated with BEACOPPescalated. On the other hand,
the improved negative predictive value of a negative PET-2 in this group of patients has to be
compared to those initially treated with ABVD.
Of note, post-hoc analyses of the UK RATHL
trial showed very good inter-observer comparability, highlighting the feasibility of utilizing
PET-2-guided therapy in routine clinical care
[125] .
1
2
3
4
5
Time (years)
Fig. 8.3 PFS in 260 patients with advanced-stage disease
according to the IPS and PET-2 status after 2×ABVD
(Reprinted with permission from Gallamini et al. [121])
Prognostic Indices
in Advanced-Stage Hodgkin
Lymphoma
Prognostic indices for advanced HL are clinically
important to select patients prior to initiation of
first-line treatment who may be over- or undertreated by standard therapeutic approaches.
Over the last decades, several groups developed prognostic indices or scores based on a
few hundred cases trying to define high-risk
groups. Most scores combined age, disease
extent, and laboratory markers, but only a prognostic model for OS developed by Gobbi et al.
based on the following seven factors received
P. J. Bröckelmann and L. Specht
154
PET-2 negative
PET-2 positive
Non-randomised
100
Progression-free surival (%)
90
80
70
60
3-year estimate (95% CI)
50
PET-2 negative
40
5-year estimate (95% CI)
93·5% (91·9–95·1)
91·4% (89·5–93·4)
92·5% (90·7–94·3)
88·3% (85·8–90·8)
30
Non-randomised 77·6% (68·6–86·6)
77·6% (68·6–86·6)
20
Overall
92·3% (91·1-93·5)
89·4% (87·9-91·0)
10
Median observation time 52 months (IQR 32–67)
PET-2 positive
0
0
Number at risk
(number censored)
PET-2 negative 1010 (0)
PET-2 positive 948 (0)
Non-randomised 115 (0)
12
24
36
48
60
921 (63)
854 (62)
72 (31)
845 (118)
766 (131)
61 (37)
685 (265)
600 (284)
47 (49)
548 (396)
476 (398)
35 (61)
385 (553)
276 (588)
24 (72)
Fig. 8.4 PFS according to PET-2 status after 2×BEACOPPescalated in 2073 patients with advanced-stage treated within
the HD18 trial (Reprinted with permission from Borchmann et al. [123])
8×eBEACOPP or 6xeBEACOPP
4×eBEACOPP
100
Progression-free surival (%)
90
80
70
60
3-year estimate (95% CI)
50
40
30
8×eBEACOPP or 6×eBEACOPP
91·7% (89·0 to 94·4)
90·8% (87·9 to 93·7)
4×eBEACOPP
95·3% (93·3 to 97·3)
92·2% (89·4 to 95·0)
3·6% (−0·2 to 6·9)
−1·4% (−2·7 to 5·4)
Difference
20
Hazard ratio (95% CI)
10
5-year estimate (95% CI)
0·79% (0·50 to 1·24)
Median observation time 55 months (IQR 36–70)
0
0
12
24
36
48
60
411 (21)
444 (22)
378 (45)
412 (43)
310 (102)
334 (119)
247 (164)
269 (180)
164 (246)
198 (246)
Number at risk
(number
censored)
8×eBEACOPP
or 6×eBEACOPP 446 (0)
4×eBEACOPP 474 (0)
Fig. 8.5 PFS in 920 PET-2-negative patients treated with a total of 6–8× vs. 4×BEACOPPescalated in the HD18 trial
(Reprinted with permission from Borchmann et al. [123])
8
Prognostic Factors
general acceptance: stage, age, histology, B
symptoms, serum albumin, sex, and involved
area distribution (infradiaphragmatic disease or
>3 supradiaphragmatic areas) [101].
Some years later, the International Prognostic
Factors Project on advanced-stage HL focussed
on FFP [92]. Individual patient data were collected from 23 centres and study groups including 5141 advanced-stage HL patients who were
treated mainly with doxorubicin-containing chemotherapy with and without radiotherapy. A
prognostic score was developed from this data set
in patients up to 65 years of age. The score is a
simple count of seven binary adverse prognostic
factors (Table 8.3):
This prognostic model, termed the
“International Prognostic Score” (IPS), predicts
5-year tumour control rates in the range of
45–80%. Each additional factor reduces the
prognosis by about 8%. Figure 8.6 shows the
FFP according to the number of adverse prognostic factors.
Since its publication, the IPS has performed
reasonably well in independent data sets [126–
131], including patients treated with intensified
BEACOPP chemotherapy, where outcome uniformly improved in all IPS groups with persisting
but quantitatively reduced differences [126, 130].
Comparison of several prognostic models [94,
128] revealed that none of the models including
the IPS are able to select either a very low-risk
group (e.g. <10% failure rate) or a substantial
very high-risk group (>50%). The prognostic
models thus discriminate only between low-risk
and high-risk patients (e.g. IPS ≤2 vs. IPS >2). A
more recent analysis verified the prognostic value
Table 8.3 Adverse prognostic factors incorporated in the
International Prognostic Factors Project score for FFP in
advanced-stage HL
Age ≥ 45 years
Male sex
Stage IV disease
Haemoglobin <10.5 g/dl
Serum albumin <4.0 g/dl
Leucocytosis ≥15 × 109/l
Lymphocytopenia <0.6 × 109/l or <8% of white blood
cell count
155
of the original IPS-7 in patients treated with either
ABVD or Stanford V within the multicentre US
E2496 trial, though with narrowed prognostic
range. Based on the three prognostic factors age,
stage, and haemoglobin identified by multivariable analysis, a simple IPS-3 with better discrimination for freedom from progression and overall
survival (OS) was developed (Fig. 8.7) [132].
Several authors tried to extend the IPS beyond
advanced stages. The IPS works nicely to predict
outcome after autologous haematopoietic stem
cell transplantation [133]. It appears to be moderately predictive in early and intermediate stages,
extending the factor stage IV to include any
extranodal disease [57, 134].
8.6
Prognostic Factors
in Relapsed or Refractory
Hodgkin Lymphoma
8.6.1
atients Treated for r/r HL
P
with Conventional Treatment
Patients relapsing after initial treatment with chemotherapy only or combined modality therapy,
whether for limited or advanced disease, historically had a poor prognosis with conventional
chemotherapy, obtaining durable remissions in
only 10–30% of cases [135–140]. In this setting,
the extent and duration of an initial remission
(i.e. time-to-relapse, TTR) is the most important
prognostic factor for outcome after relapse.
Patients who never achieve a remission (i.e.
refractory or primary progressive disease) have
an extremely poor prognosis, while patients who
relapse within or after 12 months of completion
of therapy have an intermediate or relatively
good prognosis, respectively [135–138, 140,
141]. But even for the latter, long-term outlook is
poor with historic conventional chemotherapy
and dismal in patients with second or higher
relapse [142–144]. Figure 8.8 shows survival
curves for patients relapsing after initial chemotherapy divided into these three prognostic
groups [145].
In addition to TTR, the extent of disease at
relapse is also independently significant for
P. J. Bröckelmann and L. Specht
156
1.0
# of factors
.9
% freedom from progression
Fig. 8.6 Freedom from
progression according to
the number of adverse
prognostic factors (see
Table 8.3) for 1618
patients with advanced-­
stage HL in the
International Prognostic
Factors Project
(Reprinted with
permission from
Hasenclever and Diehl
[81])
.8
0 : N=115 (7%)
1 : N=360 (22%)
.7
2 : N=464 (29%)
.6
3 : N=378 (23%)
.5
4 : N=190 (12%)
5+: N=111 (7%)
.4
.3
.2
.1
0.0
0
1
2
3
4
5
6
7
years
OS by PS-3
0.6
0.6
0.8
0.8
1.0
1.0
FFP by PS-3
OS
Log rank test P < 0.0001
0.4
0.4
FFP
Log rank test P = 0.0001
N
83 ± 2
74 ± 3
68 ± 5
63 ± 10
PS-3 = 0
PS-3 = 1
PS-3 = 2
PS-3 = 3
5-year OS (%)
474
275
82
23
95 ± 1
85 ± 2
75 ± 5
52 ± 10
0.0
0.0
0.2
5-year FFP (%)
474
275
82
23
0.2
N
PS-3 = 0
PS-3 = 1
PS-3 = 2
PS-3 = 3
0
2
4
6
8
10
12
Years
0
2
4
6
8
10
12
Years
Fig. 8.7 FFP and OS according to IPS-3 among 854 patients with advanced-stage HL in the E2496 trial (Reprinted
with permission from Diefenbach et al. [132])
prognosis. Advanced stage, extranodal disease,
and more than three involved sites at relapse are
adverse prognostic factors [100, 135, 136, 140,
146]. Age, performance status, histology other
than nodular sclerosis, B symptoms at relapse,
and a low haemoglobin have also been shown to
be significant [100, 135, 137, 138, 140, 141,
146]. Prognostic factors which have been shown
to be independently significant for outcome after
HL relapse treated with conventional chemotherapy are summarised in Table 8.4.
A subgroup of patients with relapsed or
refractory HL (r/r HL) have anatomically limited
disease. For selected patients in this subgroup,
RT with or without additional chemotherapy
offers some chance of durable remission [138,
147–150]. Prognostic factor analyses indicate
that patients suitable for this kind of relapse
treatment are those relapsing exclusively in
supradiaphragmatic nodal sites, with no B symptoms at relapse, and after a disease-free interval
of >12 months [147, 148, 151]. In patients with
8
Prognostic Factors
157
1.0
Probability
0.8
0.6
late relapse 52/169
early relapse 64/138
0.4
primary progressive 129/206
0.2
p < 0,0001
0.0
0
12
24
36
48
60
Months
72
84
96
108
120
Fig. 8.8 OS of patients with primary progressive, early relapse, or late relapse of HL, treated in GHSG trials from 1988
to 1999 primarily with conventional salvage (Reprinted with permission from Josting and Schmitz [145])
Table 8.4 Independent prognostic factors for outcome in
r/r HL treated with conventional chemotherapy
Extent and durability of first remission (TTR)
Extent of disease at relapse (relapse stage, extranodal
relapse, ≥3 sites of relapse)
B symptoms at relapse
Haemoglobin at relapse
Histology
Age
Performance status
these favourable characteristics, durable remission with RT may be achieved in up to 50% of
cases.
8.6.2
Treatment with High-Dose
Chemotherapy
and Autologous Stem Cell
Transplantation
High-dose chemotherapy (HDCT) with autologous stem cell transplantation (ASCT) is superior
to conventional chemotherapy in r/r HL [152,
153]. Hence, it is the preferred treatment in
relapsed eligible patients and current standard of
care. Similar to treatment with conventional chemotherapy, a number of prognostic factors were
identified decades ago and are still independently
significant for outcome. The chemosensitivity of
the disease is extremely important. Thus, the
response to initial or salvage therapy, the duration
of initial remission, and the number of prior
failed regimens have been shown to be associated
with outcome [154–161]. Evaluation of response
to salvage treatment before HDCT by PET further refines prognostication [162–167].
The disease burden before transplantation is
another important prognostic factor, and measures reflecting tumour burden such as stage of
disease and bulky or extranodal disease at salvage have been shown to be independently significant [145, 154, 161, 168]. B symptoms, low
haemoglobin, and elevated serum LDH at relapse
are also significant [154, 157, 162, 169, 170]. A
poor performance status is an important adverse
prognostic feature [155, 159, 171], whereas age
has not been significant in most series, probably
due to the fact that most patients are relatively
young as required by the intensive HDCT + ASCT
approach [161, 172–176]. Of note, paediatric
patients have the same outcome as adults [177].
P. J. Bröckelmann and L. Specht
158
Table 8.5 Independent prognostic factors for outcome
after HDCT + ASCT in r/r HL
Chemosensitivity of HL
Response to initial or salvage therapy
Duration of initial remission (TTR)
Number of failed prior regimens
FDG-PET after salvage and before HCT + ASCT
Disease burden before salvage
Stage of disease in r/r HL
Bulky disease in r/r HL
Extranodal r/r HL
B symptoms at r/r HL
Haemoglobin at r/r HL
Serum lactate dehydrogenase (LDH) at r/r HL
The seven factors included in the IPS for
advanced-stage HL have also been examined in
r/r HL and are at least partially applicable [133,
178]. The prognostic factors known to be independently significant for outcome after high-dose
chemotherapy and stem cell transplantation are
shown in Table 8.5.
For patients with disease recurrence after
ASCT, prognosis was poor historically.
Refractory disease at second-line treatment and
short disease-free interval after ASCT are very
poor prognostic factors [179–182].
8.6.3
atients’ Prognostic Indices or
P
Scores in r/r HL Treated
with Novel Agents
Over the last years, several studies aimed at combining individual risk factors in prognostic scores
developed by multivariable analyses to improve
clinical applicability.
An early prognostic score based on the equally
weighted three clinical risk factors haemoglobin
<10.5 g/dl and <12.0 g/dl in female and male
patients, respectively, clinical stage III/IV, and
TTR <12 months was established by an analysis
of 422 patients relapsing after treatment in GHSG
trials between 1988 and 1999. Four-year FF2F
was estimated at 65%, 35%, and 25% for low-,
intermediate-, and high-risk patients, respectively [183]. In a more recent analysis of the
European HDR2 trial, only stage IV disease but
not stage III or III/IV at relapse retained prognostic significance; 3-year PFS was estimated at
81% and 14% for the most favourable and unfavourable risk groups with zero or all three risk
factors, respectively [184].
Based on EBMT registry data and the four
factors Karnofsky performance score < 90%,
chemotherapy resistance prior to ASCT, ≥3 chemotherapies pre-ASCT, and extranodal disease, a
different risk score was proposed and externally
validated. Four-year PFS for the low-, intermediate-, and high-risk group were 71%, 60%, and
42%, respectively [185].
In a very recent international effort, a simple
externally validated prognostic score in contemporarily treated patients within or outside clinical
trials was developed [171]. Equal weighting of
the five factors stage IV disease, TTR ≤ 3 months,
ECOG performance status ≥ 1, bulk ≥ 5 cm, and
inadequate response to salvage chemotherapy
(i.e. <PR or PET positivity) thereby resulted in
optimal prognostication of both PFS and OS after
ASCT (Fig. 8.9).
8.6.4
Prognostic Factors
for Treatment with Novel
Agents in r/r HL
The anti-CD30 antibody-drug conjugate brentuximab vedotin (BV) was one of the first drugs in
decades approved for the treatment of r/r HL
based on acceptable toxicity and high response
rates in the pivotal phase II trial [186]. Subgroup
analyses according to potential risk factors did
not reveal any unfavourable risk groups in terms
of response rates or PFS. A recent trial investigating BV as salvage therapy prior to ASCT identified MTV in addition to refractory disease as
relevant prognostic factors for event-free survival
(EFS) in multivariable analysis [50]. When
administering BV as consolidative therapy after
ASCT, a sustained 5-year PFS benefit compared
to placebo was seen across risk groups with a
pronounced difference in patients with any ≥2
risk factors from a variety of potential single risk
factors [187].
8
Prognostic Factors
159
1.0
0.9
0.8
PFS rate
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0.0
0
12
1
36
24
2
3-5
48
60
72
84
Time [months]
1.0
Overall survival [months] rate
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0.0
0
12
24
1
36
2
3-5
48
60
72
84
Time [months]
Fig. 8.9 PFS and OS after ASCT for r/r HL according to the number of risk factors present (Adapted with permission
from Bröckelmann et al. 2017 [171])
More recently, the anti-PD1 antibodies
nivolumab and pembrolizumab were approved
for r/r cHL due to high response rates and exceptional tolerability in the pivotal phase II trials
[188, 189]. Across both trials, also heavily pretreated patients with unfavourable disease characteristics responded to therapy and achieved
relatively long PFS. Exploratory analyses in r/r
cHL patients treated with nivolumab revealed
9p24.1 alterations in virtually all patients and a
correlation of PFS with the 9p24.1 genetic categories (e.g. amplification or copy gain) as well as
PD-L1 expression by the HRS cells. In addition,
sustained MHC class II expression on HRS cells
was associated with superior response rates and
PFS in patients treated >12 months after
ASCT. As expected, patients achieving a PR or
CR had superior PFS to non-responders [190].
8.7
Conclusion
and Future Aspects
As demonstrated above, a large number of variables have been shown to possess prognostic significance in HL, both at initial presentation and at
relapse or progression. Most of these variables
are correlated with the total tumour burden,
P. J. Bröckelmann and L. Specht
160
which has hitherto been too cumbersome to measure for use in clinical practice. However, the
metabolic tumour volume, measured by FDG-­
PET, is highly correlated with the total tumour
burden and may become an important and clinically useful prognostic tool in the future. Today,
treatment is tailored to risk groups defined by
prognostic factors, with less intensive therapies
for patients with favourable disease in order to
reduce toxicity and increasing treatment intensity
for patients with unfavourable characteristics
with the aim of increasing cure rates. Different
centres and groups use slightly differing criteria
for initial treatment selection, which makes direct
comparisons challenging and some form of international harmonisation desirable. Early interim
functional imaging by FDG-PET as a marker for
treatment sensitivity allows individualized treatment in routine clinical care of advanced-stage
HL. In early-stage disease, its role is less clear.
Molecular abnormalities in either tumour cells or
their microenvironment as well as circulating
cell-free DNA may potentially prove to be powerful prognostic markers. Integration of these
biomarkers will hopefully enable refined therapeutic approaches better tailored to individual
disease activity and risk of treatment failure.
With the advent of novel agents, drug-specific
prognostic or predictive factors are of increasing
interest to stratify treatments and minimize
potentially harmful and expensive exposure.
7.
8.
9.
10.
11.
12.
13.
14.
15.
References
1. Reed DM (1902) On the pathological changes in
Hodgkin's disease, with especial reference to its
relation to tuberculosis. Johns Hopkins Hosp Rep
10:133–196
2. Banfi A, Bonadonna G, Buraggi G et al (1965)
Proposta di classificazione e terapia della malattia di
Hodgkin. Tumori 51:97–112
3. Easson EC, Russell MH (1963) The cure of
Hodgkin’s disease. BMJ 1963:1704–1707
4. Jelliffe AM, Thomson AD (1955) The prognosis in
Hodgkin’s disease. Br J Cancer 9:21–36
5. Kaplan HS (1970) On the natural history, treatment
and prognosis of Hodgkin’s disease. Harvey lectures
1968–1969. Academic, New York, NY, pp 215–259
6. Musshoff K, Stamm H, Lummel G, Gössel K (1964)
Zur prognose der lymphogranulomatose. Klinisches
16.
17.
18.
19.
bild un strahlentherapie. Freiburger Krankengut
1938–1958. In: Keiderling W (ed) Beiträge zur
Inneren Medizin. FK Schattauer-Verlag, Stuttgart,
pp 549–561
Peters MV (1950) A study of survivals in Hodgkin’s
disease treated radiologically. Am J Roentgenol
63:299–311
Rosenberg SA (1966) Report of the committee
on the staging of Hodgkin’s disease. Cancer Res
26:1310–1310
Carbone PP, Kaplan HS, Musshoff K, Smithers
DW, Tubiana M (1971) Report of the committee on
Hodgkin’s disease staging classification. Cancer Res
31:1860–1861
Aisenberg AC, Qazi R (1976) Improved survival in
Hodgkin’s disease. Cancer 37:2423–2429
Davis S, Dahlberg S, Myers MH, Chen A, Steinhorn
SC (1987) Hodgkin's disease in the United States:
a comparison of patient characteristics and survival
in the centralized cancer patient data system and the
surveillance, epidemiology, and end results program.
J Natl Cancer Inst 78:471–478
Gobbi PG, Cavalli C, Federico M, Bertoloni D, Di
Prisco UA, Rossi A, Silingardi V, Mauri C, Ascari E
(1988) Hodgkin’s disease prognosis: a directly predictive equation. Lancet 1:675–679
Hancock BW, Aitken M, Martin JF, Dunsmore IR,
Ross CM, Carr I, Emmanuel IG (1979) Hodgkin’s
disease in Sheffield (1971–76) (with computer analysis of variables). Clin Oncol 5:283–297
Henry-Amar M, Aeppli DM, Anderson J, Ashley S,
Bonichon F, Cox RS, Dahlberg SJ, DeBoer G, Dixon
DO, Gobbi PG, Gregory W, Hasenclever D, Löffler
M, Pompe Kirn V, Santarelli MT, Specht L, Swindell
R, Vaughan Hudson B (1990) Workshop statistical
report. In: Somers R, Henry-Amar M, Meerwaldt
JH, Carde P (eds) Treatment strategy in Hodgkin’s
disease. INSERM/John Libbey Eurotext, London,
pp 169–425
Kaplan HS (1973) Survival and relapse rates in
Hodgkin’s disease: Stanford experience, 1961–71.
Monogr Natl Cancer Inst 36:487–496
Musshoff K, Hartmann C, Niklaus B, Rössnere
R (1974) Results of therapy in Hodgkin’s disease:
Freiburg i. Br. 1964–1971. In: Musshoff K (ed)
Diagnosis and therapy of malignant lymphoma.
Springer, Berlin, pp 206–220
Sutcliffe SB, Gospodarowicz MK, Bergsagel DE,
Bush RS, Alison RE, Bean HA, Brown TC, Chua T,
Clark RM, Curtis JE (1985) Prognostic groups for
management of localized Hodgkin’s disease. J Clin
Oncol 3:393–401
Noone AM, Howlader N, Krapcho M, Miller D,
Brest A, Yu M, Ruhl J, Tatalovich Z, Mariotto A,
Lewis DR, Chen HS, Feuer EJ, Cronin KA (2018)
SEER cancer statistics review, 1975–2015. National
Cancer Institute, Bethesda, MD. https://seer.cancer.
gov/csr/1975_2015
Lister TA, Crowther D, Sutcliffe SB, Glatstein E,
Canellos GP, Young RC, Rosenberg SA, Coltman
8
Prognostic Factors
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
CA, Tubiana M (1989) Report of a committee
convened to discuss the evaluation and staging of
patients with Hodgkin’s disease: Cotswolds meeting. J Clin Oncol 7:1630–1636
Cheson BD, Fisher RI, Barrington SF, Cavalli
F, Schwartz LH, Zucca E, Lister TA (2014)
Recommendations for initial evaluation, staging, and
response assessment of Hodgkin and non-Hodgkin
lymphoma: the Lugano classification. J Clin Oncol
32:3059–3068
George SL (1988) Identification and assessment of
prognostic factors. Semin Oncol 15:462–471
Gospodarowicz MK, O'Sullivan B, Koh ES (2006)
Prognostic factors: principles and applications. In:
Gospodarowicz MK, O’Sullivan B, Sobin LH (eds)
Prognostic factors in cancer, 3rd edn. Wiley-Liss,
Hoboken, NJ, pp 23–38
Byar DP (1988) Identification of prognostic factors. In: Buyse ME, Staquet MJ, Sylvester RJ (eds)
Cancer clinical trials. Methods and practice. Oxford
University Press, Oxford, pp 423–443
Riley RD, Sauerbrei W, Altman DG (2009)
Prognostic markers in cancer: the evolution of evidence from single studies to meta-analysis, and
beyond. Br J Cancer 100:1219–1229
Byar DP (1991) Problems with using observational databases to compare treatments. Stat Med
10:663–666
Simon R (1984) Importance of prognostic factors in
cancer clinical trials. Cancer Treat Rep 68:185–192
Cheson BD, Pfistner B, Juweid ME, Gascoyne
RD, Specht L, Horning SJ, Coiffier B, Fisher RI,
Hagenbeek A, Zucca E, Rosen ST, Stroobants S,
Lister TA, Hoppe RT, Dreyling M, Tobinai K, Vose
JM, Connors JM, Federico M, Diehl V (2007)
Revised response criteria for malignant lymphoma.
J Clin Oncol 25:579–586
Altman DG (2006) Studies investigating prognostic
factors: conduct and evaluation. In: Gospodarowicz
MK, O’Sullivan B, Sobin LH (eds) Prognostic factors in cancer, 3rd edn. Wiley-Liss, Hoboken, NJ,
pp 39–54
Burke HB, Henson DE (1993) The American Joint
Committee on Cancer. Criteria for prognostic factors and for an enhanced prognostic system. Cancer
72:3131–3135
Cox DR (1972) Regression models and life-tables. J
R Stat Soc B 34:187–220
Simon R (2001) Evaluating prognostic factor studies. In: Gospodarowicz MK, Henson DE, RVP
H, O’Sullivan B, Sobin LH, Wittekind C (eds)
Prognostic factors in cancer, 2nd edn. Wiley-Liss,
New York, NY, pp 49–56
Specht L, Nordentoft AM, Cold S, Clausen NT,
Nissen NI (1987) Tumour burden in early stage
Hodgkin’s disease: the single most important prognostic factor for outcome after radiotherapy. Br J
Cancer 55:535–539
Specht L, Nissen NI (1988) Prognostic factors
in Hodgkin’s disease stage IV. Eur J Haematol
41:359–367
161
34. Specht L, Nissen NI (1988) Prognostic factors in
Hodgkin’s disease stage III with special reference to
tumour burden. Eur J Haematol 41:80–87
35. Specht L, Nissen NI (1988) Hodgkin’s disease stages
I and II with infradiaphragmatic presentation: a rare
and prognostically unfavourable combination. Eur J
Haematol 40:396–402
36. Specht L, Nordentoft AM, Cold S, Clausen NT,
Nissen NI (1988) Tumor burden as the most important
prognostic factor in early stage Hodgkin’s disease.
Relations to other prognostic factors and implications
for choice of treatment. Cancer 61:1719–1727
37. Specht L, Nissen NI (1989) Hodgkin's disease and
age. Eur J Haematol 43:127–135
38. Specht L, Lauritzen AF, Nordentoft AM, Andersen
PK, Christensen BE, Hippe E, Hou-Jensen K, Nissen
NI (1990) Tumor cell concentration and tumor burden in relation to histopathologic subtype and other
prognostic factors in early stage Hodgkin's disease.
The Danish National Hodgkin Study Group. Cancer
65:2594–2601
39. Specht L (1992) Tumour burden as the main indicator of prognosis in Hodgkin’s disease. Eur J Cancer
28A:1982–1985
40. Gobbi PG, Ghirardelli ML, Solcia M, Di Giulio G,
Merli F, Tavecchia L, Berte R, Davini O, Levis A,
Broglia C, Maffe GC, Ilariucci F, Dore R, Ascari
E (2001) Image-aided estimate of tumor burden in
Hodgkin’s disease: evidence of its primary prognostic importance. J Clin Oncol 19:1388–1394
41. Gobbi PG, Broglia C, Di Giulio G, Mantelli M,
Anselmo P, Merli F, Zinzani PL, Rossi G, Callea
V, Iannitto E, Paulli M, Garioni L, Ascari E (2004)
The clinical value of tumor burden at diagnosis in
Hodgkin lymphoma. Cancer 101:1824–1834
42. Gobbi PG, Valentino F, Bassi E, Coriani C, Merli
F, Bonfante V, Marchiano A, Gallamini A, Bolis
S, Stelitano C, Levis A, Federico M, Angrilli F,
Di GG, Corazza GR (2011) Chemoresistance as a
function of the pretherapy tumor burden and the
chemotherapy regimen administered: differences
observed with 2 current chemotherapy regimens
for advanced Hodgkin lymphoma. Clin Lymphoma
Myeloma Leuk 11:396–402
43. Gobbi PG, Bergonzi M, Bassi E, Merli F, Coriani C,
Stelitano C, Iannitto E, Federico M (2012) Tumor
burden in Hodgkin’s lymphoma can be reliably estimated from a few staging parameters. Oncol Rep
28:815–820
44. Gobbi PG, Bassi E, Bergonzi M, Merli F, Coriani
C, Iannitto E, Luminari S, Polimeno G, Federico M
(2012) Tumour burden predicts treatment resistance
in patients with early unfavourable or advanced
stage Hodgkin lymphoma treated with ABVD and
radiotherapy. Hematol Oncol 30:194–199
45. Gobbi PG, Bergonzi M, Bassi E, Merli F, Coriani
C, Federico M (2013) Tumour burden at diagnosis
as the main clinical predictor of cell resistance in
patients with early stage, favourable Hodgkin lymphoma treated with VBM chemotherapy plus radiotherapy. Hematol Oncol 31:151–155
P. J. Bröckelmann and L. Specht
162
46. Gobbi PG (2014) Tumor burden in Hodgkin's lymphoma: much more than the best prognostic factor.
Crit Rev Oncol Hematol 90:17–23
47. Akhtari M, Milgrom SA, Pinnix CC, Reddy JP,
Dong W, Smith GL, Mawlawi O, Abou YZ, Gunther
J, Osborne EM, Andraos TY, Wogan CF, Rohren E,
Garg N, Chuang H, Khoury JD, Oki Y, Fanale M,
Dabaja BS (2018) Reclassifying patients with early-­
stage Hodgkin lymphoma based on functional radiographic markers at presentation. Blood 131:84–94
48. Cottereau AS, Versari A, Loft A, Casasnovas O,
Bellei M, Ricci R, Bardet S, Castagnoli A, Brice P,
Raemaekers J, Deau B, Fortpied C, Raveloarivahy
T, Van ZE, Chartier L, Vander BT, Federico M,
Hutchings M, Ricardi U, Andre M, Meignan M
(2018) Prognostic value of baseline metabolic tumor
volume in early-stage Hodgkin lymphoma in the
standard arm of the H10 trial. Blood 131:1456–1463
49. Kanoun S, Rossi C, Berriolo-Riedinger A, Dygai-­
Cochet I, Cochet A, Humbert O, Toubeau M, Ferrant
E, Brunotte F, Casasnovas RO (2014) Baseline metabolic tumour volume is an independent prognostic
factor in Hodgkin lymphoma. Eur J Nucl Med Mol
Imaging 41:1735–1743
50. Moskowitz AJ, Schoder H, Gavane S, Thoren KL,
Fleisher M, Yahalom J, McCall SJ, Cadzin BR, Fox
SY, Gerecitano J, Grewal R, Hamlin PA, Horwitz
SM, Kumar A, Matasar M, Ni A, Noy A, Palomba
ML, Perales MA, Portlock CS, Sauter C, Straus
D, Younes A, Zelenetz AD, Moskowitz CH (2017)
Prognostic significance of baseline metabolic tumor
volume in relapsed and refractory Hodgkin lymphoma. Blood 130:2196–2203
51. Prochazka V, Gawande RS, Cayci Z, Froelich
JW, Cao Q, Wilke C, Dusenbery K, Weisdorf DJ,
Bachanova V (2018) Positron emission tomography-­
based assessment of metabolic tumor volume predicts survival after autologous hematopoietic cell
transplantation for Hodgkin lymphoma. Biol Blood
Marrow Transplant 24:64–70
52. Gallamini A, Kostakoglu L (2017) Metabolic tumor
volume: we still need a platinum-standard metric. J
Nucl Med 58:196–197
53. Ilyas H, Mikhaeel NG, Dunn JT, Rahman F, Moller
H, Smith D, Barrington SF (2018) Defining the
optimal method for measuring baseline metabolic
tumour volume in diffuse large B cell lymphoma.
Eur J Nucl Med Mol Imaging 45:1142–1154
54. Kanoun S, Tal I, Berriolo-Riedinger A, Rossi C,
Riedinger JM, Vrigneaud JM, Legrand L, Humbert
O, Casasnovas O, Brunotte F, Cochet A (2015)
Influence of software tool and methodological
aspects of Total metabolic tumor volume calculation on baseline [18F]FDG PET to predict survival
in Hodgkin lymphoma. PLoS One 10:e0140830
55. Kostakoglu L, Chauvie S (2018) Metabolic tumor
volume metrics in lymphoma. Semin Nucl Med
48:50–66
56. Bonfante V, Santoro A, Viviani S, Zucali R, Devizzi
L, Zanini M, Tess JDT, Valagussa P, Banfi A,
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
Bonadonna G (1993) Early stage Hodgkin’s disease:
ten-year results of a non-randomised study with
radiotherapy alone or combined with MOPP. Eur J
Cancer 29A:24–29
Franklin J, Paulus U, Lieberz D, Breuer K, Tesch H,
Diehl V (2000) Is the international prognostic score
for advanced stage Hodgkin’s disease applicable to
early stage patients? German Hodgkin Lymphoma
Study Group. Ann Oncol 11:617–623
Glimelius I, Molin D, Amini RM, Gustavsson A,
Glimelius B, Enblad G (2003) Bulky disease is the
most important prognostic factor in Hodgkin lymphoma stage IIB. Eur J Haematol 71:327–333
Gobbi PG, Gendarini A, Crema A, Cavalli C,
Attardo-Parrinello G, Federico M, Di Prisco U,
Ascari E (1985) Serum albumin in Hodgkin’s disease. Cancer 55:389–393
Gospodarowicz MK, Sutcliffe SB, Clark RM,
Dembo AJ, Fitzpatrick PJ, Munro AJ, Bergsagel DE,
Patterson BJ, Tsang R, Chua T (1992) Analysis of
supradiaphragmatic clinical stage I and II Hodgkin’s
disease treated with radiation alone. Int J Radiat
Oncol Biol Phys 22:859–865
Pavlovsky S, Maschio M, Santarelli MT, Muriel
FS, Corrado C, Garcia I, Schwartz L, Montero C,
Sanahuja FL, Magnasco O (1988) Randomized trial
of chemotherapy versus chemotherapy plus radiotherapy for stage I–II Hodgkin’s disease. J Natl
Cancer Inst 80:1466–1473
Specht L (1991) Prognostic factors in Hodgkin’s disease. Cancer Treat Rev 18:21–53
Tubiana M, Henry-Amar M, Werf-Messing B,
Henry J, Abbatucci J, Burgers M, Hayat M, Somers
R, Laugier A, Carde P (1985) A multivariate analysis
of prognostic factors in early stage Hodgkin’s disease. Int J Radiat Oncol Biol Phys 11:23–30
Tubiana M, Henry-Amar M, Carde P, Burgers JM,
Hayat M, van der Schueren E, Noordijk EM, Tanguy
A, Meerwaldt JH, Thomas J (1989) Toward comprehensive management tailored to prognostic factors
of patients with clinical stages I and II in Hodgkin’s
disease. The EORTC lymphoma group controlled
clinical trials: 1964–1987. Blood 73:47–56
Vaughan Hudson B, MacLennan KA, Bennett MH,
Easterling MJ, Vaughan Hudson G, Jelliffe AM
(1987) Systemic disturbance in Hodgkin’s disease
and its relation to histopathology and prognosis
(BNLI report no. 30). Clin Radiol 38:257–261
Filippi AR, Botticella A, Bello M, Botto B,
Castiglione A, Gavarotti P, Gottardi D, Parvis G, Bisi
G, Levis A, Vitolo U, Ricardi U (2013) Interim positron emission tomography and clinical outcome in
patients with early stage Hodgkin lymphoma treated
with combined modality therapy. Leuk Lymphoma
54:1183–1187
Gallamini A, Kostakoglu L (2012) Interim FDG-­
PET in Hodgkin lymphoma: a compass for a safe
navigation in clinical trials? Blood 120:4913–4920
Hutchings M, Loft A, Hansen M, Pedersen LM,
Buhl T, Jurlander J, Buus S, Keiding S, D'Amore
8
Prognostic Factors
69.
70.
71.
72.
73.
74.
75.
76.
77.
78.
F, Boesen AM, Berthelsen AK, Specht L (2006)
FDG-PET after two cycles of chemotherapy predicts
treatment failure and progression-free survival in
Hodgkin lymphoma. Blood 107:52–59
Hutchings M, Kostakoglu L, Zaucha JM, Malkowski
B, Biggi A, Danielewicz I, Loft A, Specht L,
Lamonica D, Czuczman MS, Nanni C, Zinzani
PL, Diehl L, Stern R, Coleman M (2014) In vivo
treatment sensitivity testing with positron emission
tomography/computed tomography after one cycle
of chemotherapy for Hodgkin lymphoma. J Clin
Oncol 32:2705–2711
Kostakoglu L, Schoder H, Johnson JL, Hall NC,
Schwartz LH, Straus DJ, LaCasce AS, Jung SH,
Bartlett NL, Canellos GP, Cheson BD (2012)
Interim [(18)F]fluorodeoxyglucose positron emission tomography imaging in stage I-II non-bulky
Hodgkin lymphoma: would using combined positron emission tomography and computed tomography criteria better predict response than each test
alone? Leuk Lymphoma 53:2143–2150
Zinzani PL, Rigacci L, Stefoni V, Broccoli A,
Puccini B, Castagnoli A, Vaggelli L, Zanoni L,
Argnani L, Baccarani M, Fanti S (2012) Early
interim 18F-FDG PET in Hodgkin’s lymphoma:
evaluation on 304 patients. Eur J Nucl Med Mol
Imaging 39:4–12
Sher DJ, Mauch PM, Van den Abbeele A, LaCasce
AS, Czerminski J, Ng AK (2009) Prognostic significance of mid- and post-ABVD PET imaging in
Hodgkin’s lymphoma: the importance of involved-­
field radiotherapy. Ann Oncol 20:1848–1853
Adams HJ, Nievelstein RA, Kwee TC (2015)
Prognostic value of interim FDG-PET in Hodgkin
lymphoma: systematic review and meta-analysis. Br
J Haematol 170:356–366
Adams HJ, Nievelstein RA, Kwee TC (2016)
Systematic review and meta-analysis on the prognostic value of complete remission status at FDG-­
PET in Hodgkin lymphoma after completion of
first-line therapy. Ann Hematol 95:1–9
Brockelmann
PJ,
Angelopoulou
MK,
Vassilakopoulos TP (2016) Prognostic factors in
Hodgkin lymphoma. Semin Hematol 53:155–164
Alonso-Alvarez S, Vidriales MB, Caballero MD,
Blanco O, Puig N, Martin A, Penarrubia MJ, Zato
E, Galende J, Barez A, Alcoceba M, Orfao A,
Gonzalez M, Garcia-Sanz R (2017) The number of
tumor infiltrating T-cell subsets in lymph nodes from
patients with Hodgkin lymphoma is associated with
the outcome after first line ABVD therapy. Leuk
Lymphoma 58:1144–1152
Calio A, Zamo A, Ponzoni M, Zanolin ME, Ferreri
AJ, Pedron S, Montagna L, Parolini C, Fraifeld
VE, Wolfson M, Yanai H, Pizzolo G, Doglioni C,
Vinante F, Chilosi M (2015) Cellular senescence
markers p16INK4a and p21CIP1/WAF are predictors of Hodgkin lymphoma outcome. Clin Cancer
Res 21:5164–5172
Casasnovas RO, Mounier N, Brice P, Divine M,
Morschhauser F, Gabarre J, Blay JY, Voillat L,
163
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
Lederlin P, Stamatoullas A, Bienvenu J, Guiguet M,
Intrator L, Grandjean M, Briere J, Ferme C, Salles
G (2007) Plasma cytokine and soluble receptor
signature predicts outcome of patients with classical Hodgkin’s lymphoma: a study from the Groupe
d’Etude des Lymphomes de l’Adulte. J Clin Oncol
25:1732–1740
Derenzini E, Younes A (2011) Predicting treatment
outcome in classical Hodgkin lymphoma: genomic
advances. Genome Med 3:26
Diefenbach C, Steidl C (2013) New strategies in
Hodgkin lymphoma: better risk profiling and novel
treatments. Clin Cancer Res 19:2797–2803
Guo B, Cen H, Tan X, Ke Q (2016) Meta-analysis of
the prognostic and clinical value of tumor-associated
macrophages in adult classical Hodgkin lymphoma.
BMC Med 14:159
Guo X, Wang J, Jin J, Chen H, Zhen Z, Jiang W, Lin
T, Huang H, Xia Z, Sun X (2018) High serum level
of soluble programmed death ligand 1 is associated
with a poor prognosis in Hodgkin lymphoma. Transl
Oncol 11:779–785
Hagenbeek A, Gascoyne RD, Dreyling M, Kluin P,
Engert A, Salles G (2009) Biomarkers and prognosis
in malignant lymphomas. Clin Lymphoma Myeloma
9:160–166
Kanakry JA, Li H, Gellert LL, Lemas MV, Hsieh
WS, Hong F, Tan KL, Gascoyne RD, Gordon LI,
Fisher RI, Bartlett NL, Stiff P, Cheson BD, Advani
R, Miller TP, Kahl BS, Horning SJ, Ambinder RF
(2013) Plasma Epstein-Barr virus DNA predicts outcome in advanced Hodgkin lymphoma: correlative
analysis from a large north American cooperative
group trial. Blood 121:3547–3553
Koh YW, Kang HJ, Park C, Yoon DH, Kim S, Suh C,
Go H, Kim JE, Kim CW, Huh J (2012) The ratio of
the absolute lymphocyte count to the absolute monocyte count is associated with prognosis in Hodgkin’s
lymphoma: correlation with tumor-associated macrophages. Oncologist 17:871–880
Mestre F, Gutierrez A, Ramos R, Martinez-Serra J,
Sanchez L, Matheu G, Ros T, Garcia JF, Rodriguez
J (2012) Expression of COX-2 on Reed-Sternberg
cells is an independent unfavorable prognostic factor
in Hodgkin lymphoma treated with ABVD. Blood
119:6072–6079
Navarro A, Munoz C, Gaya A, Diaz-Beya M, Gel
B, Tejero R, Diaz T, Martinez A, Monzo M (2013)
MiR-SNPs as markers of toxicity and clinical outcome in Hodgkin lymphoma patients. PLoS One
8:e64716
Scott DW, Chan FC, Hong F, Rogic S, Tan KL,
Meissner B, Ben-Neriah S, Boyle M, Kridel R,
Telenius A, Woolcock BW, Farinha P, Fisher RI,
Rimsza LM, Bartlett NL, Cheson BD, Shepherd
LE, Advani RH, Connors JM, Kahl BS, Gordon
LI, Horning SJ, Steidl C, Gascoyne RD (2013)
Gene expression-based model using formalin-fixed
paraffin-­embedded biopsies predicts overall survival
in advanced-stage classical Hodgkin lymphoma. J
Clin Oncol 31:692–700
P. J. Bröckelmann and L. Specht
164
89. Steidl C, Lee T, Shah SP, Farinha P, Han G, Nayar
T, Delaney A, Jones SJ, Iqbal J, Weisenburger DD,
Bast MA, Rosenwald A, Muller-Hermelink HK,
Rimsza LM, Campo E, Delabie J, Braziel RM,
Cook JR, Tubbs RR, Jaffe ES, Lenz G, Connors JM,
Staudt LM, Chan WC, Gascoyne RD (2010) Tumor-­
associated macrophages and survival in classic
Hodgkin’s lymphoma. N Engl J Med 362:875–885
90. Touati M, Delage-Corre M, Monteil J, Abraham J,
Moreau S, Remenieras L, Gourin MP, Dmytruk N,
Olivrie A, Turlure P, Girault S, Labrousse F, Preux
PM, Jaccard A, Bordessoule D (2015) CD68-­positive
tumor-associated macrophages predict unfavorable
treatment outcomes in classical Hodgkin lymphoma
in correlation with interim fluorodeoxyglucose-­
positron emission tomography assessment. Leuk
Lymphoma 56:332–341
91. Vassilakopoulos
TP,
Dimopoulou
MN,
Angelopoulou MK, Petevi K, Pangalis GA,
Moschogiannis M, Dimou M, Boutsikas G,
Kanellopoulos A, Gainaru G, Plata E, Flevari P,
Koutsi K, Papageorgiou L, Telonis V, Tsaftaridis
P, Sachanas S, Yiakoumis X, Tsirkinidis P, Viniou
NA, Siakantaris MP, Variami E, Kyrtsonis MC,
Meletis J, Panayiotidis P, Konstantopoulos K (2016)
Prognostic implication of the absolute lymphocyte
to absolute monocyte count ratio in patients with
classical Hodgkin lymphoma treated with doxorubicin, bleomycin, vinblastine, and Dacarbazine or
equivalent regimens. Oncologist 21:343–353
92. Hasenclever D, Diehl V (1998) A prognostic score
for advanced Hodgkin’s disease. International prognostic factors project on advanced Hodgkin’s disease. N Engl J Med 339:1506–1514
93. Canellos GP, Anderson JR, Propert KJ, Nissen
N, Cooper MR, Henderson ES, Green MR,
Gottlieb A, Peterson BA (1992) Chemotherapy of
advanced Hodgkin’s disease with MOPP, ABVD,
or MOPP alternating with ABVD. N Engl J Med
327:1478–1484
94. Ferme C, Bastion Y, Brice P, Lederlin P, Divine M,
Gabarre J, Assouline D, Ferrant A, Berger F, Lepage
E (1997) Prognosis of patients with advanced
Hodgkin’s disease: evaluation of four prognostic
models using 344 patients included in the group
d’Etudes des Lymphomes de l’Adulte study. Cancer
80:1124–1133
95. Peterson BA, Pajak TF, Cooper MR, Nissen NI,
Glidewell OJ, Holland JF, Bloomfield CD, Gottlieb
AJ (1982) Effect of age on therapeutic response and
survival in advanced Hodgkin’s disease. Cancer
Treat Rep 66:889–898
96. Sjoberg J, Halthur C, Kristinsson SY, Landgren O,
Nygell UA, Dickman PW, Bjorkholm M (2012)
Progress in Hodgkin lymphoma: a population-based
study on patients diagnosed in Sweden from 1973–
2009. Blood 119:990–996
97. Straus DJ, Gaynor JJ, Myers J, Merke DP, Caravelli
J, Chapman D, Yahalom J, Clarkson BD (1990)
Prognostic factors among 185 adults with newly
98.
99.
100.
101.
102.
103.
104.
105.
106.
107.
diagnosed advanced Hodgkin’s disease treated with
alternating potentially noncross- resistant chemotherapy and intermediate-dose radiation therapy. J
Clin Oncol 8:1173–1186
Proctor SJ, Rueffer JU, Angus B, Breuer K, Flechtner
H, Jarrett R, Levis A, Taylor P, Tirelli U (2002)
Hodgkin’s disease in the elderly: current status and
future directions. Ann Oncol 13(Suppl 1):133–137
Engert A, Ballova V, Haverkamp H, Pfistner B,
Josting A, Duhmke E, Muller-Hermelink K, Diehl
V (2005) Hodgkin’s lymphoma in elderly patients:
a comprehensive retrospective analysis from the
German Hodgkin’s study group. J Clin Oncol
23:5052–5060
Boll B, Goergen H, Arndt N, Meissner J, Krause SW,
Schnell R, von TB, Eichenauer DA, Sasse S, Fuchs
M, Behringer K, Klimm BC, Naumann R, Diehl V,
Engert A, Borchmann P (2013) Relapsed hodgkin
lymphoma in older patients: a comprehensive analysis from the German Hodgkin Study Group. J Clin
Oncol 31:4431–4437
Gobbi PG, Comelli M, Grignani GE, Pieresca C,
Bertoloni D, Ascari E (1994) Estimate of expected
survival at diagnosis in Hodgkin’s disease: a means
of weighting prognostic factors and a tool for treatment choice and clinical research. A report from
the international database on Hodgkin’s disease
(IDHD). Haematologica 79:241–255
Klimm B, Reineke T, Haverkamp H, Behringer K,
Eich HT, Josting A, Pfistner B, Diehl V, Engert A
(2005) Role of hematotoxicity and sex in patients
with Hodgkin’s lymphoma: an analysis from
the German Hodgkin study group. J Clin Oncol
23:8003–8011
Torricelli P, Grimaldi PL, Fiocchi F, Federico M,
Romagnoli R (2004) Hodgkin’s disease: a quantitative evaluation by computed tomography of tumor
burden. Clin Imaging 28:239–244
Axdorph U, Sjoberg J, Grimfors G, Landgren
O, Porwit-MacDonald A, Bjorkholm M (2000)
Biological markers may add to prediction of o­ utcome
achieved by the international prognostic score in
Hodgkin’s disease. Ann Oncol 11:1405–1411
Nadali G, Vinante F, Ambrosetti A, Todeschini G,
Veneri D, Zanotti R, Meneghini V, Ricetti MM,
Benedetti F, Vassanelli A (1994) Serum levels of soluble CD30 are elevated in the majority of untreated
patients with Hodgkin’s disease and correlate
with clinical features and prognosis. J Clin Oncol
12:793–797
Zanotti R, Trolese A, Ambrosetti A, Nadali G, Visco
C, Ricetti MM, Benedetti F, Pizzolo G (2002) Serum
levels of soluble CD30 improve international prognostic score in predicting the outcome of advanced
Hodgkin’s lymphoma. Ann Oncol 13:1908–1914
Sauer M, Plutschow A, Jachimowicz RD, Kleefisch
D, Reiners KS, Ponader S, Engert A, von Strandmann
EP (2013) Baseline serum TARC levels predict therapy outcome in patients with Hodgkin lymphoma.
Am J Hematol 88:113–115
8
Prognostic Factors
108. Sureda A, Connors JM, Younes A, Gallamini A,
Ansell SM, Kim WS, Miall F, Bajel A, Knopinska-­
Posluszny W, Ogden CA, Kuroda S, Liu R,
Trepicchio WL, Radford J (2018) Serum sCD30 and
TARC do not correlate with PETZ-based response
assessment in patients (pts) with stage III or IV
classical Hodgkin lymphoma (cHL). HemaSphere
2:35–36
109. Bohlen H, Kessler M, Sextro M, Diehl V, Tesch
H (2000) Poor clinical outcome of patients with
Hodgkin’s disease and elevated interleukin-10
serum levels. Clinical significance of interleukin-10
serum levels for Hodgkin’s disease. Ann Hematol
79:110–113
110. Marri PR, Hodge LS, Maurer MJ, Ziesmer SC,
Slager SL, Habermann TM, Link BK, Cerhan
JR, Novak AJ, Ansell SM (2013) Prognostic significance of pretreatment serum cytokines in
classical Hodgkin lymphoma. Clin Cancer Res
19:6812–6819
111. Vassilakopoulos TP, Nadali G, Angelopoulou MK,
Siakantaris MP, Dimopoulou MN, Kontopidou
FN, Karkantaris C, Kokoris SI, Kyrtsonis MC,
Tsaftaridis P, Pizzolo G, Pangalis GA (2002) The
prognostic significance of beta(2)-microglobulin in
patients with Hodgkin’s lymphoma. Haematologica
87:701–708
112. Meador CB, Lovly CM (2015) Liquid biopsies
reveal the dynamic nature of resistance mechanisms
in solid tumors. Nat Med 21:663–665
113. Vandenberghe P, Wlodarska I, Tousseyn T, Dehaspe
L, Dierickx D, Verheecke M, Uyttebroeck A,
Bechter O, Delforge M, Vandecaveye V, Brison
N, Verhoef GE, Legius E, Amant F, Vermeesch JR
(2015) Non-invasive detection of genomic imbalances in Hodgkin/Reed-Sternberg cells in early and
advanced stage Hodgkin’s lymphoma by sequencing
of circulating cell-free DNA: a technical proof-of-­
principle study. Lancet Haematol 2:e55–e65
114. Bräuninger A, Desch A-K, Kunze K, Boten A,
Brobeil A, Rummel M, Kurch L, Georgi T, Kluge
R, Mauz-Körholz C, Körholz D, Gattenlöhner
S (2018) Genotyping circulating tumor DNA of
pediatric Hodgkin lymphoma patients to determine pathogenic mechanisms and monitor therapy.
HemaSphere 2:4–4
115. Jin MC, Kurtz DM, Esfahani MS, Sworder BJ,
Schroer-Martin J, Soo J, Glover C, Roschewski
M, Wilson W, Düehrsen U, Hüttmann A, Rossi D,
Gaidano G, Westin J, Diehn M, Advani R, Alizadeh
AA (2018) Circulating tumor DNA as a biomarker
for the noninvasive genotyping and monitoring of
classical Hodgkin lymphoma. HemaSphere 2:4–5
116. Tan KL, Scott DW, Hong F, Kahl BS, Fisher RI,
Bartlett NL, Advani RH, Buckstein R, Rimsza
LM, Connors JM, Steidl C, Gordon LI, Horning
SJ, Gascoyne RD (2012) Tumor-associated macrophages predict inferior outcomes in classic Hodgkin
lymphoma: a correlative study from the E2496 intergroup trial. Blood 120:3280–3287
165
117. Roemer MG, Advani RH, Ligon AH, Natkunam Y,
Redd RA, Homer H, Connelly CF, Sun HH, Daadi
SE, Freeman GJ, Armand P, Chapuy B et al (2016)
PD-L1 and PD-L2 genetic alterations define classical Hodgkin lymphoma and predict outcome. J Clin
Oncol 34:2690–2697
118. Gallamini A, Rigacci L, Merli F, Nassi L, Bosi A,
Capodanno I, Luminari S, Vitolo U, Sancetta R,
Iannitto E, Trentin L, Stelitano C, Tavera S, Biggi
A, Castagnoli A, Versari A, Gregianin M, Pelosi E,
Torchio P, Levis A (2006) The predictive value of
positron emission tomography scanning performed
after two courses of standard therapy on treatment
outcome in advanced stage Hodgkin's disease.
Haematologica 91:475–481
119. Hutchings M, Mikhaeel NG, Fields PA, Nunan T,
Timothy AR (2005) Prognostic value of interim
FDG-PET after two or three cycles of chemotherapy
in Hodgkin lymphoma. Ann Oncol 16:1160–1168
120. Zinzani PL, Tani M, Fanti S, Alinari L, Musuraca G,
Marchi E, Stefoni V, Castellucci P, Fina M, Farshad
M, Pileri S, Baccarani M (2006) Early positron
emission tomography (PET) restaging: a predictive
final response in Hodgkin’s disease patients. Ann
Oncol 17:1296–1300
121. Gallamini A, Hutchings M, Rigacci L, Specht L,
Merli F, Hansen M, Patti C, Loft A, Di Raimondo F,
D'Amore F, Biggi A, Vitolo U, Stelitano C, Sancetta
R, Trentin L, Luminari S, Iannitto E, Viviani S, Pierri
I, Levis A (2007) Early interim 2-[18F]fluoro-­
2-­
deoxy-D-glucose positron emission tomography is
prognostically superior to international prognostic
score in advanced-stage Hodgkin’s lymphoma: a
report from a joint Italian-Danish study. J Clin Oncol
25:3746–3752
122. Johnson P, Federico M, Kirkwood A, Fossa A,
Berkahn L, Carella A, D'Amore F, Enblad G,
Franceschetto A, Fulham M, Luminari S, O'Doherty
M, Patrick P, Roberts T, Sidra G, Stevens L, Smith P,
Trotman J, Viney Z, Radford J, Barrington S (2016)
Adapted treatment guided by interim PET-CT scan
in advanced Hodgkin’s lymphoma. N Engl J Med
374:2419–2429
123. Borchmann P, Goergen H, Kobe C, Lohri A, Greil R,
Eichenauer DA, Zijlstra JM, Markova J, Meissner J,
Feuring-Buske M, Huttmann A, Dierlamm J, Soekler
M, Beck HJ, Willenbacher W, Ludwig WD, Pabst T,
Topp MS, Hitz F, Bentz M, Keller UB, Kuhnhardt D,
Ostermann H, Schmitz N, Hertenstein B, Aulitzky
W, Maschmeyer G, Vieler T, Eich H, Baues C, Stein
H, Fuchs M, Kuhnert G, Diehl V, Dietlein M, Engert
A (2018) PET-guided treatment in patients with
advanced-stage Hodgkin’s lymphoma (HD18): final
results of an open-label, international, randomised
phase 3 trial by the German Hodgkin Study Group.
Lancet 390:2790–2802
124. Casanovas O, Brice P, Bouabdallah R, Salles G,
Stamatoullas A, Dupuis J, Reman O, Gastinne T,
Joly B, Bouabdallah K, Nicolas-Virelizier E, Feugier
P, Morschhauser F, Delarue R, Berriolo-­Riedinger A,
P. J. Bröckelmann and L. Specht
166
125.
126.
127.
128.
129.
130.
131.
Edeline V, Traverse-Giehen A, Andre M, Mounier N,
Meignan M (2018) Final analysis of the AHL2011
randomized phase III LYSA study comparing an
early PET driven treatment de-­escalation to a not
PET-monitored strategy in patients with advanced
stages Hodgkin lymphoma. EHA Learning Center,
The Hague. https://learningcenter.ehaweb.org/
eha/2018/stockholm/214504/olivier.casasnovas.
final.analysis.of.the.ahl2011.randomized.phase.iii.
lysa.html?f=topic=1574∗media=3
Barrington SF, Kirkwood AA, Franceschetto A,
Fulham MJ, Roberts TH, Almquist H, Brun E,
Hjorthaug K, Viney ZN, Pike LC, Federico M,
Luminari S, Radford J, Trotman J, Fossa A, Berkahn
L, Molin D, D'Amore F, Sinclair DA, Smith P,
O'Doherty MJ, Stevens L, Johnson PW (2016)
PET-CT for staging and early response: results from
the response-adapted therapy in advanced Hodgkin
lymphoma study. Blood 127:1531–1538
Diehl V, Franklin J, Pfreundschuh M, Lathan B,
Paulus U, Hasenclever D, Tesch H, Herrmann R,
Dorken B, Muller-Hermelink HK, Duhmke E,
Loeffler M (2003) Standard and increased-dose
BEACOPP chemotherapy compared with COPP-­
ABVD for advanced Hodgkin’s disease. N Engl J
Med 348:2386–2395
Duggan DB, Petroni GR, Johnson JL, Glick JH,
Fisher RI, Connors JM, Canellos GP, Peterson BA
(2003) Randomized comparison of ABVD and
MOPP/ABV hybrid for the treatment of advanced
Hodgkin’s disease: report of an intergroup trial. J
Clin Oncol 21:607–614
Gobbi PG, Zinzani PL, Broglia C, Comelli M,
Magagnoli M, Federico M, Merli F, Iannitto E,
Tura S, Ascari E (2001) Comparison of prognostic
models in patients with advanced Hodgkin disease.
Promising results from integration of the best three
systems. Cancer 91:1467–1478
Johnson PW, Radford JA, Cullen MH, Sydes MR,
Walewski J, Jack AS, MacLennan KA, Stenning SP,
Clawson S, Smith P, Ryder D, Hancock BW (2005)
Comparison of ABVD and alternating or hybrid
multidrug regimens for the treatment of advanced
Hodgkin's lymphoma: results of the United Kingdom
lymphoma group LY09 trial (ISRCTN97144519). J
Clin Oncol 23:9208–9218
Moccia AA, Donaldson J, Chhanabhai M, Hoskins
PJ, Klasa RJ, Savage KJ, Shenkier TN, Slack GW,
Skinnider B, Gascoyne RD, Connors JM, Sehn LH
(2012) International prognostic score in advanced-­
stage Hodgkin’s lymphoma: altered utility in the
modern era. J Clin Oncol 30:3383–3388
Radford JA, Rohatiner AZ, Ryder WD, Deakin DP,
Barbui T, Lucie NP, Rossi A, Dunlop DJ, Cowan
RA, Wilkinson PM, Gupta RK, James RD, Shamash
J, Chang J, Crowther D, Lister TA (2002) ChlVPP/
EVA hybrid versus the weekly VAPEC-B regimen
for previously untreated Hodgkin’s disease. J Clin
Oncol 20:2988–2994
132. Diefenbach CS, Li H, Hong F, Gordon LI, Fisher
RI, Bartlett NL, Crump M, Gascoyne RD, Wagner
H Jr, Stiff PJ, Cheson BD, Stewart DA, Kahl BS,
Friedberg JW, Blum KA, Habermann TM, Tuscano
JM, Hoppe RT, Horning SJ, Advani RH (2015)
Evaluation of the international prognostic score
(IPS-7) and a simpler prognostic score (IPS-3) for
advanced Hodgkin lymphoma in the modern era. Br
J Haematol 171:530–538
133. Bierman PJ, Lynch JC, Bociek RG, Whalen VL,
Kessinger A, Vose JM, Armitage JO (2002) The
international prognostic factors project score for
advanced Hodgkin’s disease is useful for predicting outcome of autologous hematopoietic stem cell
transplantation. Ann Oncol 13:1370–1377
134. Gisselbrecht C, Mounier N, Andre M, Casanovas
O, Reman O, Sebban C, Divine M, Brice P, Briere
J, Hennequin C, Ferme C (2005) How to define
intermediate stage in Hodgkin’s lymphoma? Eur J
Haematol Suppl 75:111–114
135. Bonfante V, Santoro A, Viviani S, Devizzi L,
Balzarotti M, Soncini F, Zanini M, Valagussa P,
Bonadonna G (1997) Outcome of patients with
Hodgkin’s disease failing after primary MOPP-­
ABVD. J Clin Oncol 15:528–534
136. Brice P, Bastion Y, Divine M, Nedellec G, Ferrant
A, Gabarre J, Reman O, Lepage E, Ferme C (1996)
Analysis of prognostic factors after the first relapse
of Hodgkin’s disease in 187 patients. Cancer
78:1293–1299
137. Ferme C, Bastion Y, Lepage E, Berger F, Brice P,
Morel P, Gabarre J, Nedellec G, Reman O, Cheron
N (1995) The MINE regimen as intensive salvage
chemotherapy for relapsed and refractory Hodgkin’s
disease. Ann Oncol 6:543–549
138. Lohri A, Barnett M, Fairey RN, O'Reilly SE, Phillips
GL, Reece D, Voss N, Connors JM (1991) Outcome
of treatment of first relapse of Hodgkin’s disease
after primary chemotherapy: identification of risk
factors from the British Columbia experience 1970
to 1988. Blood 77:2292–2298
139. Longo DL, Duffey PL, Young RC, Hubbard SM,
Ihde DC, Glatstein E, Phares JC, Jaffe ES, Urba
WJ, DeVita VT Jr (1992) Conventional-dose salvage
combination chemotherapy in patients relapsing
with Hodgkin’s disease after combination chemotherapy: the low probability for cure. J Clin Oncol
10:210–218
140. Viviani S, Santoro A, Negretti E, Bonfante V,
Valagussa P, Bonadonna G (1990) Salvage chemotherapy in Hodgkin’s disease. Results in patients
relapsing more than twelve months after first complete remission. Ann Oncol 1:123–127
141. Josting A, Franklin J, May M, Koch P, Beykirch
MK, Heinz J, Rudolph C, Diehl V, Engert A (2001)
New prognostic score based on treatment outcome
of patients with relapsed Hodgkin’s lymphoma registered in the database of the German Hodgkin’s
Lymphoma Study Group. J Clin Oncol 20:221–230
8
Prognostic Factors
142. Hagemeister FB, Tannir N, McLaughlin P, Salvador
P, Riggs S, Velasquez WS, Cabanillas F (1987)
MIME chemotherapy (methyl-GAG, ifosfamide,
methotrexate, etoposide) as treatment for recurrent
Hodgkin’s disease. J Clin Oncol 5:556–561
143. Perren TJ, Selby PJ, Milan S, Meldrum M,
McElwain TJ (1990) Etoposide and adriamycin containing combination chemotherapy (HOPE-Bleo) for
relapsed Hodgkin’s disease. Br J Cancer 61:919–923
144. Straus DJ, Myers J, Koziner B, Lee BJ, Clarkson BD
(1983) Combination chemotherapy for the treatment
of Hodgkin’s disease in relapse. Results with lomustine (CCNU), melphalan (Alkeran), and vindesine
(DVA) alone (CAD) and in alternation with MOPP
and doxorubicin (Adriamycin), bleomycin, and
vinblastine (ABV). Cancer Chemother Pharmacol
11:80–85
145. Josting A, Schmitz N (2004) Insights into 25 years
of clinical trials of the GHSG: Relapsed and refractory Hodgkin’s disease. In: Diehl V, Josting A (eds)
25 years German Hodgkin Study Group. Urban &
Vogel, Munich, pp 89–99
146. Salvagno L, Soraru M, Aversa SM, Bianco A,
Chiarion S, Pappagallo GL, Fiorentino MV (1993)
Late relapses in Hodgkin’s disease: outcome of
patients relapsing more than twelve months after
primary chemotherapy. Ann Oncol 4:657–662
147. Brada M, Eeles R, Ashley S, Nichols J, Horwich A
(1992) Salvage radiotherapy in recurrent Hodgkin’s
disease. Ann Oncol 3:131–135
148. Josting A, Nogova L, Franklin J, Glossmann JP, Eich
HT, Sieber M, Schober T, Boettcher HD, Schulz U,
Muller RP, Diehl V, Engert A (2005) Salvage radiotherapy in patients with relapsed and refractory
Hodgkin’s lymphoma: a retrospective analysis from
the German Hodgkin lymphoma study group. J Clin
Oncol 23:1522–1529
149. Leigh BR, Fox KA, Mack CF, Baier M, Miller TP,
Cassady JR (1993) Radiation therapy salvage of
Hodgkin’s disease following chemotherapy failure.
Int J Radiat Oncol Biol Phys 27:855–862
150. Roach M III, Kapp DS, Rosenberg SA, Hoppe RT
(1987) Radiotherapy with curative intent: an option
in selected patients relapsing after chemotherapy
for advanced Hodgkin’s disease. J Clin Oncol
5:550–555
151. Wirth A, Corry J, Laidlaw C, Matthews J, Liew KH
(1997) Salvage radiotherapy for Hodgkin’s disease
following chemotherapy failure. Int J Radiat Oncol
Biol Phys 39:599–607
152. Linch DC, Winfield D, Goldstone AH, Moir D,
Hancock B, McMillan A, Chopra R, Milligan D,
Hudson GV (1993) Dose intensification with autologous bone-marrow transplantation in relapsed and
resistant Hodgkin’s disease: results of a BNLI randomised trial. Lancet 341:1051–1054
153. Schmitz N, Pfistner B, Sextro M, Sieber M, Carella
AM, Haenel M, Boissevain F, Zschaber R, Muller P,
Kirchner H, Lohri A, Decker S, Koch B, Hasenclever
167
154.
155.
156.
157.
158.
159.
160.
D, Goldstone AH, Diehl V (2002) Aggressive
conventional chemotherapy compared with high-­
dose chemotherapy with autologous haemopoietic
stem-cell transplantation for relapsed chemosensitive Hodgkin’s disease: a randomised trial. Lancet
359:2065–2071
Argiris A, Seropian S, Cooper DL (2000) High-dose
BEAM chemotherapy with autologous peripheral
blood progenitor-cell transplantation for unselected
patients with primary refractory or relapsed
Hodgkin’s disease. Ann Oncol 11:665–672
Bierman PJ, Anderson JR, Freeman MB, Vose JM,
Kessinger A, Bishop MR, Armitage JO (1996)
High-dose chemotherapy followed by autologous
hematopoietic rescue for Hodgkin’s disease patients
following first relapse after chemotherapy. Ann
Oncol 7:151–156
Czyz A, Lojko-Dankowska A, Dytfeld D, Nowicki
A, Gil L, Matuszak M, Kozlowska-Skrzypczak
M, Kazmierczak M, Bembnista E, Komarnicki M
(2013) Prognostic factors and long-term outcome of
autologous haematopoietic stem cell transplantation
following a uniform-modified BEAM-conditioning
regimen for patients with refractory or relapsed
Hodgkin lymphoma: a single-center experience.
Med Oncol 30:611
Ferme C, Mounier N, Divine M, Brice P, Stamatoullas
A, Reman O, Voillat L, Jaubert J, Lederlin P, Colin
P, Berger F, Salles G (2002) Intensive salvage therapy with high-dose chemotherapy for patients with
advanced Hodgkin’s disease in relapse or failure
after initial chemotherapy: results of the Groupe
d’Etudes des Lymphomes de l’Adulte H89 trial. J
Clin Oncol 20:467–475
Fernandez de Larrea C, Martinez C, Gaya A,
Lopez-­Guillermo A, Rovira M, Fernandez-Aviles
F, Lozano M, Bosch F, Esteve J, Nomdedeu B,
Montserrat E, Carreras E (2010) Salvage chemotherapy with alternating MINE-ESHAP regimen
in relapsed or refractory Hodgkin’s lymphoma followed by autologous stem-cell transplantation. Ann
Oncol 21:1211–1216
Hahn T, Benekli M, Wong C, Moysich KB, Hyland
A, Michalek AM, Alam A, Baer MR, Bambach B,
Czuczman MS, Wetzler M, Becker JL, McCarthy
PL (2005) A prognostic model for prolonged event-­
free survival after autologous or allogeneic blood
or marrow transplantation for relapsed and refractory Hodgkin’s disease. Bone Marrow Transplant
35:557–566
Sureda A, Constans M, Iriondo A, Arranz R,
Caballero MD, Vidal MJ, Petit J, Lopez A, Lahuerta
JJ, Carreras E, Garcia-Conde J, Garcia-Larana J,
Cabrera R, Jarque I, Carrera D, Garcia-Ruiz JC,
Pascual MJ, Rifon J, Moraleda JM, Perez-Equiza
K, Albo C, Diaz-Mediavilla J, Torres A, Torres
P, Besalduch J, Marin J, Mateos MV, Fernandez-­
Ranada JM, Sierra J, Conde E (2005) Prognostic
factors affecting long-term outcome after stem cell
P. J. Bröckelmann and L. Specht
168
161.
162.
163.
164.
165.
166.
167.
168.
transplantation in Hodgkin’s lymphoma autografted
after a first relapse. Ann Oncol 16:625–633
Viviani S, Di NM, Bonfante V, Di SA, Carlo-Stella
C, Matteucci P, Magni M, Devizzi L, Valagussa P,
Gianni AM (2010) Long-term results of high-dose
chemotherapy with autologous bone marrow or
peripheral stem cell transplant as first salvage treatment for relapsed or refractory Hodgkin lymphoma:
a single institution experience. Leuk Lymphoma
51:1251–1259
Jabbour E, Hosing C, Ayers G, Nunez R, Anderlini P,
Pro B, Khouri I, Younes A, Hagemeister F, Kwak L,
Fayad L (2007) Pretransplant positive positron emission tomography/gallium scans predict poor outcome in patients with recurrent/refractory Hodgkin
lymphoma. Cancer 109:2481–2489
Mocikova H, Pytlik R, Markova J, Steinerova K,
Kral Z, Belada D, Trnkova M, Trneny M, Koza
V, Mayer J, Zak P, Kozak T (2011) Pre-transplant
positron emission tomography in patients with
relapsed Hodgkin lymphoma. Leuk Lymphoma
52:1668–1674
Moskowitz AJ, Yahalom J, Kewalramani T,
Maragulia JC, Vanak JM, Zelenetz AD, Moskowitz
CH (2010) Pretransplantation functional imaging
predicts outcome following autologous stem cell
transplantation for relapsed and refractory Hodgkin
lymphoma. Blood 116:4934–4937
Moskowitz CH, Yahalom J, Zelenetz AD, Zhang
Z, Filippa D, Teruya-Feldstein J, Kewalramani
T, Moskowitz AJ, Rice RD, Maragulia J, Vanak J,
Trippett T, Hamlin P, Horowitz S, Noy A, O'Connor
OA, Portlock C, Straus D, Nimer SD (2010) High-­
dose chemo-radiotherapy for relapsed or refractory Hodgkin lymphoma and the significance of
pre-transplant functional imaging. Br J Haematol
148:890–897
Moskowitz CH, Matasar MJ, Zelenetz AD, Nimer
SD, Gerecitano J, Hamlin P, Horwitz S, Moskowitz
AJ, Noy A, Palomba L, Perales MA, Portlock C,
Straus D, Maragulia JC, Schoder H, Yahalom J
(2012) Normalization of pre-ASCT, FDG-PET
imaging with second-line, non-cross-resistant,
chemotherapy programs improves event-free survival in patients with Hodgkin lymphoma. Blood
119:1665–1670
Sucak GT, Ozkurt ZN, Suyani E, Yasar DG, Akdemir
OU, Aki Z, Yegin ZA, Yagci M, Kapucu OL (2011)
Early post-transplantation positron emission tomography in patients with Hodgkin lymphoma is an
independent prognostic factor with an impact on
overall survival. Ann Hematol 90:1329–1336
Popat U, Hosing C, Saliba RM, Anderlini P, van
Besien K, Przepiorka D, Khouri IF, Gajewski J,
Claxton D, Giralt S, Rodriguez M, Romaguera J,
Hagemeister F, Ha C, Cox J, Cabanillas F, Andersson
BS, Champlin RE (2004) Prognostic factors for disease progression after high-dose chemotherapy and
autologous hematopoietic stem cell transplantation
169.
170.
171.
172.
173.
174.
175.
176.
for recurrent or refractory Hodgkin’s lymphoma.
Bone Marrow Transplant 33:1015–1023
Josting A, Rudolph C, Mapara M, Glossmann JP,
Sienawski M, Sieber M, Kirchner HH, Dorken B,
Hossfeld DK, Kisro J, Metzner B, Berdel WE, Diehl
V, Engert A (2005) Cologne high-dose sequential
chemotherapy in relapsed and refractory Hodgkin
lymphoma: results of a large multicenter study of the
German Hodgkin lymphoma study group (GHSG).
Ann Oncol 16:116–123
Lumley MA, Milligan DW, Knechtli CJ, Long SG,
Billingham LJ, McDonald DF (1996) High lactate
dehydrogenase level is associated with an adverse
outlook in autografting for Hodgkin’s disease. Bone
Marrow Transplant 17:383–388
Brockelmann PJ, Muller H, Casasnovas O,
Hutchings M, von Tresckow B, Jurgens M, McCall
SJ, Morschhauser F, Fuchs M, Borchmann P,
Moskowitz CH, Engert A (2017) Risk factors and
a prognostic score for survival after autologous
stem-cell transplantation for relapsed or refractory
Hodgkin lymphoma. Ann Oncol 28:1352–1358
Bierman PJ, Bagin RG, Jagannath S, Vose JM,
Spitzer G, Kessinger A, Dicke KA, Armitage JO
(1993) High dose chemotherapy followed by autologous hematopoietic rescue in Hodgkin’s disease:
long-term follow-up in 128 patients. Ann Oncol
4:767–773
Brice P, Bouabdallah R, Moreau P, Divine M, Andre
M, Aoudjane M, Fleury J, Anglaret B, Baruchel A,
Sensebe L, Colombat P (1997) Prognostic factors
for survival after high-dose therapy and autologous
stem cell transplantation for patients with relapsing
Hodgkin’s disease: analysis of 280 patients from
the French registry. Societe Francaise de Greffe de
Moelle. Bone Marrow Transplant 20:21–26
Chopra R, McMillan AK, Linch DC, Yuklea S,
Taghipour G, Pearce R, Patterson KG, Goldstone
AH (1993) The place of high-dose BEAM therapy
and autologous bone marrow transplantation in
poor-risk Hodgkin's disease. A single-center eight-­
year study of 155 patients. Blood 81:1137–1145
Horning SJ, Chao NJ, Negrin RS, Hoppe RT, Long
GD, Hu WW, Wong RM, Brown BW, Blume KG
(1997) High-dose therapy and autologous hematopoietic progenitor cell transplantation for recurrent
or refractory Hodgkin’s disease: analysis of the
Stanford University results and prognostic indices.
Blood 89:801–813
Wheeler C, Eickhoff C, Elias A, Ibrahim J, Ayash
L, McCauley M, Mauch P, Schwartz G, Eder JP,
Mazanet R, Ferrara J, Rimm IJ, Guinan E, Bierer
B, Gilliland G, Churchill WH, Ault K, Parsons S,
Antman K, Schnipper L, Tepler I, Gaynes L, Frei E
III, Kadin M, Antin J (1997) High-dose cyclophosphamide, carmustine, and etoposide with autologous
transplantation in Hodgkin's disease: a prognostic
model for treatment outcomes. Biol Blood Marrow
Transplant 3:98–106
8
Prognostic Factors
177. Williams CD, Goldstone AH, Pearce R, Green
S, Armitage JO, Carella A, Meloni G (1993)
Autologous bone marrow transplantation for pediatric Hodgkin’s disease: a case-matched comparison
with adult patients by the European bone marrow
transplant group lymphoma registry. J Clin Oncol
11:2243–2249
178. Sirohi B, Cunningham D, Powles R, Murphy F,
Arkenau T, Norman A, Oates J, Wotherspoon A,
Horwich A (2008) Long-term outcome of autologous
stem-cell transplantation in relapsed or refractory
Hodgkin’s lymphoma. Ann Oncol 19:1312–1319
179. Arai S, Fanale M, DeVos S, Engert A, Illidge T,
Borchmann P, Younes A, Morschhauser F, McMillan
A, Horning SJ (2013) Defining a Hodgkin lymphoma
population for novel therapeutics after relapse from
autologous hematopoietic cell transplant. Leuk
Lymphoma 54:2531–2533
180. Kaloyannidis P, Voutiadou G, Baltadakis I, Tsirigotis
P, Spyridonidis A, Repousis P, Balta A, Tsimberis S,
Karakasis D, Sakellari I, Dervenoulas I, Harhalakis
N, Anagnostopoulos A (2012) Outcomes of
Hodgkin’s lymphoma patients with relapse or progression following autologous hematopoietic cell
transplantation. Biol Blood Marrow Transplant
18:451–457
181. Moskowitz AJ, Perales MA, Kewalramani T,
Yahalom J, Castro-Malaspina H, Zhang Z, Vanak
J, Zelenetz AD, Moskowitz CH (2009) Outcomes
for patients who fail high dose chemoradiotherapy
and autologous stem cell rescue for relapsed and primary refractory Hodgkin lymphoma. Br J Haematol
146:158–163
182. von Tresckow B, Muller H, Eichenauer DA,
Glossmann JP, Josting A, Boll B, Klimm B, Sasse S,
Fuchs M, Borchmann P, Engert A (2014) Outcome
and risk factors of patients with Hodgkin lymphoma
who relapse or progress after autologous stem cell
transplant. Leuk Lymphoma 55:1922–1924
183. Josting A, Franklin J, May M, Koch P, Beykirch
MK, Heinz J, Rudolph C, Diehl V, Engert A (2002)
New prognostic score based on treatment outcome
of patients with relapsed Hodgkin’s lymphoma registered in the database of the German Hodgkin’s
lymphoma study group. J Clin Oncol 20:221–230
184. Josting A, Muller H, Borchmann P, Baars JW,
Metzner B, Dohner H, Aurer I, Smardova L, Fischer
T, Niederwieser D, Schafer-Eckart K, Schmitz
N, Sureda A, Glossmann J, Diehl V, DeJong D,
Hansmann ML, Raemaekers J, Engert A (2010) Dose
169
185.
186.
187.
188.
189.
190.
intensity of chemotherapy in patients with relapsed
Hodgkin’s lymphoma. J Clin Oncol 28:5074–5080
Hahn T, McCarthy PL, Carreras J, Zhang MJ,
Lazarus HM, Laport GG, Montoto S, Hari PN
(2013) Simplified validated prognostic model for
progression-free survival after autologous transplantation for Hodgkin lymphoma. Biol Blood Marrow
Transplant 19:1740–1744
Younes A, Gopal AK, Smith SE, Ansell SM,
Rosenblatt JD, Savage KJ, Ramchandren R, Bartlett
NL, Cheson BD et al (2012) Results of a pivotal
phase II study of brentuximab vedotin for patients
with relapsed or refractory Hodgkin’s lymphoma. J
Clin Oncol 30:2183–2189
Moskowitz CH, Walewski J, Nademanee A, Masszi
T, Agura E, Holowiecki J, Abidi MH, Chen AI, Stiff
P, Viviani S, Bachanova V, Sureda A, McClendon
T, Lee C, Lisano J, Sweetenham J (2018) Five-year
PFS from the AETHERA trial of brentuximab vedotin for Hodgkin lymphoma at high risk of progression or relapse. Blood 132(25):2639–2642
Chen R, Zinzani PL, Fanale MA, Armand P,
Johnson NA, Brice P, Radford J, Ribrag V, Molin
D, Vassilakopoulos TP, Tomita A, von Tresckow B,
Shipp MA, Zhang Y, Ricart AD, Balakumaran A,
Moskowitz CH (2017) Phase II study of the efficacy and safety of Pembrolizumab for relapsed/
refractory classic Hodgkin lymphoma. J Clin Oncol
35:2125–2132
Younes A, Santoro A, Shipp M, Zinzani PL,
Timmerman JM, Ansell S, Armand P, Fanale M,
Ratanatharathorn V, Kuruvilla J, Cohen JB, Collins
G, Savage KJ, Trneny M, Kato K, Farsaci B, Parker
SM, Rodig S, Roemer MG, Ligon AH, Engert A
(2016) Nivolumab for classical Hodgkin’s lymphoma after failure of both autologous stem-cell
transplantation and brentuximab vedotin: a multicentre, multicohort, single-arm phase 2 trial. Lancet
Oncol 17:1283–1294
MGM R, Redd RA, Cader FZ, Pak CJ, Abdelrahman
S, Ouyang J, Sasse S, Younes A, Fanale M,
Santoro A, Zinzani PL, Timmerman J, Collins GP,
Ramchandren R, Cohen JB, de Boer JP, Kuruvilla
J, Savage KJ, Trneny M, Ansell S, Kato K, Farsaci
B, Sumbul A, Armand P, Neuberg DS, Pinkus GS,
Ligon AH, Rodig SJ, Shipp MA (2018) Major histocompatibility complex class II and programmed
death ligand 1 expression predict outcome after programmed death 1 blockade in classic Hodgkin lymphoma. J Clin Oncol 36:942–950
9
Principles of Radiation Therapy
for Hodgkin Lymphoma
Joachim Yahalom, Bradford S. Hoppe,
Joanna C. Yang, and Richard T. Hoppe
Contents
9.1
Principles of Radiation Therapy of Hodgkin Lymphoma 173
9.2
The Evolution of Radiotherapy for HL 173
9.3
9.3.1
9.3.2
9.3.3
9.3.4
Indications for Radiation Therapy in HL Lymphocyte-Predominant HL Classic Hodgkin: Stage I–II Stage III–IV HL RT in Salvage Programs for Refractory and Relapsed HL 174
174
174
177
178
Radiation Fields and Volumes: Principles and Design Extended-Field Radiation Therapy Involved-Field Radiation Therapy Involved-Site Radiation Therapy (ISRT): The New Standard
Volume for HL 9.4.3.1 ISRT When RT Is the Primary Treatment 9.4.3.2 ISRT When RT Is Part of Combined-Modality Treatment 9.4.4
Involved-Node Radiation Therapy (INRT): A Special Case of ISRT 9.4.5
Volume Definitions for Planning ISRT and INRT 9.4.5.1 Volume of Interest Acquisition 179
179
179
9.4
9.4.1
9.4.2
9.4.3
180
180
180
181
182
182
J. Yahalom (*)
Department of Radiation Oncology, Memorial Sloan
Kettering Cancer Center, New York, NY, USA
e-mail: yahalomj@mskcc.org
B. S. Hoppe
Department of Radiation Oncology,
Mayo Clinic Florida, Jacksonville, FL, USA
e-mail: hoppe.bradford@mayo.edu
J. C. Yang
Department of Radiation Oncology, University of
California San Francisco, San Francisco, CA, USA
e-mail: joanna.yang@ucsf.edu
R. T. Hoppe
Department of Radiation Oncology, Stanford
University Medical Center, Stanford, CA, USA
e-mail: rhoppe@stanford.edu
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_9
171
J. Yahalom et al.
172
9.4.5.2
9.4.5.3
9.4.5.4
9.4.5.5
9.4.6
9.4.6.1
9.4.6.2
9.4.6.3
9.4.6.4
9.4.7
etermination of Gross Tumor Volume (GTV) D
Determination of Clinical Target Volume (CTV) Determination of Internal Target Volume (ITV) Determination of Planning Target Volume (PTV) Determination of Organs at Risk (OAR) Lung Heart Thyroid Second Cancers Consolidation Volume Radiation Therapy (CVRT) 182
182
182
182
183
183
183
184
184
184
9.5
9.5.1
9.5.2
ose Considerations and Recommendations D
The Significance of Reducing the Radiation Dose Dose Recommendations 184
185
186
9.6
9.6.1
9.6.2
ew Aspects of Radiation Volume Definition and Treatment Delivery N
New Technologies Deep Inspiration Breath Hold 186
187
187
9.7
Proton Therapy 192
9.8
ommon Side Effects and Supportive Care During Radiation
C
Therapy Common Acute Side Effects Uncommon Early Side Effects Supportive Care During Treatment Follow-Up After Treatment 193
193
193
193
194
9.8.1
9.8.2
9.8.3
9.8.4
References 194
Abbreviations
EORTC
3DCRT
FFTF
GELA
ABVD
AP-PA
ASCT
ASH
BEACOPP
BV
CR
CT
CTV
CVRT
DIBH
EBVP
EFS
Three-dimensional
conformal
radiotherapy
Adriamycin (doxorubicin), bleomycin, vinblastine, dacarbazine
Opposed anterior and posterior
fields
Autologous
stem
cell
transplantation
American Society of Hematology
Bleomycin, etoposide, doxorubicin, cyclophosphamide, procarbazine, prednisone
Brentuximab vedotin
Complete response
Computed tomography
Clinical target volume
Consolidation volume radiation
therapy
Deep-Inspiration Breath Hold
Epirubicin, bleomycin, vinblastine, dacarbazine
Event-free survival
GHSG
HL
IFRT
IMRT
INRT
ISHL11
ISRT
LPHL
MOP-BAP
MOPP
MSKCC
MTD
European
Organisation
for
Research and Treatment of Cancer
Freedom from treatment failure
Groupe d’Études des Lymphomes
Adultes
German Hodgkin Study Group
Hodgkin lymphoma
Involved-field radiation therapy
Intensity-modulated
radiation
therapy
Involved-node radiation therapy
International
Symposium
on
Hodgkin
Lymphoma
2018
meeting
Involved-site radiation therapy
Lymphocyte-predominant HL
Mechlorethamine,
vincristine,
prednisone, bleomycin, doxorubicin, procarbazine
Mustargen, vincristine, procarbazine, prednisone
Memorial Sloan Kettering Cancer
Center
Maximum tumor dimension
9
Principles of Radiation Therapy for Hodgkin Lymphoma
NCCN
OS
PET
PTV
RT
STLI
TLI
TSH
9.1
National Comprehensive Cancer
Network
Overall survival
Positron emission tomography
Planning target volume
Radiation therapy
Subtotal lymphoid irradiation
Total lymphoid irradiation
Thyroid-stimulating hormone
Principles of Radiation
Therapy of Hodgkin
Lymphoma
173
stages of HL. In early-stage HL, multiple randomized studies have shown that the omission
of RT results in inferior progression-­free survival even after chemo-intensification.
3. Modern RT for HL treats only involved sites
to reduced doses and is both better tolerated
and associated with significantly lower risk
for long-term morbidities than the large-field,
high-dose RT used as a single modality in the
past [3].
9.2
The Evolution
of Radiotherapy for HL
RT has been used in the management of HL since
Radiation therapy (RT) is a major component of shortly after the discovery of X-rays [4, 5].
the successful treatment of Hodgkin lymphoma Initially, it was used for local palliation, but care(HL). For decades, RT was used alone to cure the ful study by pioneers in the field including Rene
majority of patients with HL, and it is still the Gilbert and Vera Peters demonstrated that more
most effective single agent in the oncologic arma- aggressive treatment with higher doses and larger
mentarium for this disease [1]. RT alone remains fields resulted in the cure of many patients, espethe treatment of choice for patients with early- cially those who presented with limited disease
stage lymphocyte-predominant HL (LPHL) and [6, 7]. At Stanford, Henry Kaplan, advantaged by
for selected patients with classic HL who have access to the medical linear accelerator, refined
contraindications to chemotherapy [2].Currently, the RT concepts and together with Saul Rosenberg
most patients with HL are treated with combined- advocated strongly for the curative potential of
modality programs in which RT is given as con- RT [8]. RT as a single modality remained the
solidation after chemotherapy. As the role of RT standard therapy for patients until effective chehas transformed over the years from a single motherapy was developed in the second half of
modality into a component of combined-­modality the twentieth century. The success of chemothertherapy, the classic principles of RT fields, dose, apy along with the awareness of adverse late
and technique have fundamentally changed.
events linked to RT initially led to a decrease in
The following principles guide the current its use, but the eventual realization that its judistrategy of using RT in HL:
cious application in lower doses and to more tailored fields could enhance curability and allow a
1. RT as part of a combined-modality program is meaningful decrease in chemotherapy doses led
radically different from the large-field, high-­ to the development of combined-modality
dose RT that was used as a single modality in programs.
the past. Both the volume treated and the dose
The RT of modern combined-modality therrequired are significantly reduced following apy programs includes the use of very limited
chemotherapy as compared to when RT was treatment volumes and the employment of
used alone. In addition, the planning and deliv- advanced techniques that improve conformity
ery of RT has improved substantially over the and dose homogeneity. In contrast to RT fields of
last two decades and continues to improve.
the past, which were based upon bony landmarks,
2. Adding RT to chemotherapy improves disease these field reductions require detailed clinical
control and allows the administration of shorter information to delineate the target accurately.
and less toxic chemotherapy regimens for all Both pre- and post-chemotherapy imaging are
J. Yahalom et al.
174
essential to define the tumor volume and the integration of computed tomography (CT), and positron emission tomography (PET)/CT treatment
planning further improves accurate RT volume
design. A margin of safety to address subclinical
disease and random and systematic positioning
error is still necessary in treatment setup, but
techniques to minimize inaccuracies in treatment
planning and delivery continue to develop.
The current recommended RT volume is
involved-site radiation therapy (ISRT), which uses
pre- and post-chemotherapy CT imaging to tailor
the radiation volumes to include only the initially
involved lymph node sites and residual CT abnormalities. ISRT represents a significant reduction
from the previous customary involved-­field RT,
which was based on bony landmarks visualized on
2D imaging. Involved-node radiation therapy
(INRT) is an even more restricted form of ISRT
and is recommended only when detailed pre-chemotherapy imaging in the treatment position is
available [9]. The volumes for ISRT and INRT
were designed to be smaller than the classic IFRT
fields that encompassed the entire predefined anatomical regions. Recommendations for ISRT and
INRT design have been established, and INRT has
already been incorporated in combined-modality
clinical trials in the European Organisation for the
Research and Treatment of Cancer (EORTC) and
the German Hodgkin Study Group (GHSG) [10].
Recommendations for ISRT design have recently
been established by the International Lymphoma
Radiation Oncology Group (ILROG), and ISRT
has been incorporated into pediatric and adult
guidelines and clinical trials in North America and
Europe [2, 11, 12].
9.3
Indications for Radiation
Therapy in HL
It is important to distinguish between classic HL
and nodular lymphocyte-predominant HL
(LPHL). The management of each entity is different. Most patients with stage I–II LPHL may
be treated with radiation alone with curative
intent, whereas combined-modality therapy is the
standard approach for the majority of patients
with classic HL.
9.3.1
Lymphocyte-Predominant HL
Over 75% of patients with LPHL present with
stage IA or IIA disease. In this setting, the disease is commonly limited to one peripheral site
(neck, axilla, or groin), and involvement of the
mediastinum is extremely rare. The National
Comprehensive Cancer Network (NCCN) guidelines [2], the German Hodgkin Lymphoma Study
Group (GHSG), and the European Organisation
for Research and Treatment of Cancer (EORTC)
currently recommend limited radiation (IFRT or
ISRT) as the treatment of choice for early-stage
LPHL. Since the mediastinum is rarely involved,
it does not need to be prophylactically treated,
thus avoiding the site most responsible for
radiation-­
related short- and long-term side
effects. In a recent retrospective study of 131
patients with stage IA disease, 98% of patients
obtained a complete response (CR), 98% after
extended-field RT alone, 100% after involved-­
field RT alone, and 95% after combined-modality
therapy [13]. With a median follow-up of
43 months, only 5% of patients relapsed and only
three patients died. Toxicity of treatment was
generally mild and was the greatest in association
with combined-modality therapy. Two other studies from the Peter MacCallum in Australia and
the Dana-Farber Cancer Institute in Boston supported the adequacy of limited-field RT for LPHL
and suggested a reduced risk of second tumors
compared to extended-field RT [14, 15].
Although there has not been a prospective
study comparing extended-field RT, which was
commonly used in the past, and IFRT/ISRT,
­retrospective data suggest that the more limited
fields are adequate [15, 16]. The radiation dose
recommended is 30–36 Gy, with the higher dose
reserved for bulky sites.
9.3.2
Classic Hodgkin: Stage I–II
Over the last two decades, the treatment of stage
I–II classic HL has changed markedly. Combined-­
modality therapy consisting of short-course chemotherapy, most often ABVD, followed by
reduced-dose IFRT/ISRT carefully directed only
to the involved lymph node(s) has replaced RT
9
Principles of Radiation Therapy for Hodgkin Lymphoma
alone as the treatment of choice. Combined
modality is the standard treatment for favorable
and unfavorable presentations of stage I–II disease in Europe, including the EORTC and
GHSG. In the United States, chemotherapy followed by ISRT is the preferred treatment recommended by the NCCN guidelines [2]. Several
randomized studies have demonstrated that
excellent results in stage I–II HL may be obtained
with combined-modality treatment that includes
only IFRT and that more extensive fields of total
or subtotal lymphoid irradiation (STLI and TLI)
are not required [17].
The strategy to reduce the number of chemotherapy cycles and/or the radiation dose was
tested by two large-scale randomized non-­
inferiority studies conducted by the GHSG. In
the HD10 study, 1370 patients with early favorable HL were randomly assigned in a 2 × 2 factorial design to receive either four or two cycles
of ABVD followed by 30 or 20 Gy IFRT. The
8-year freedom from treatment failure (FFTF)
and overall survival (OS) for all patients were
87% and 95%, respectively. Most importantly,
there were no significant differences between
patients receiving the minimal treatment of
ABVD x two cycles followed by IFRT of only
20 Gy and patients receiving more chemotherapy and/or more RT [18]. Patients with unfavorable early-­
stage HL were randomized on the
GHSG HD11 to receive either four cycles of
ABVD or four cycles of baseline BEACOPP,
followed by IFRT of either 30 or 20 Gy. Fiveyear FFTF and OS for all patients were 85% and
94.5%, respectively. There was no difference in
FFTF when BEACOPP × 4 cycles was followed
by either 30 or 20 Gy, and similar excellent
results were obtained with ABVD × 4 cycles and
IFRT of 30 Gy. Patients who received
ABVD×4 cycles and only 20 Gy had a FFTF that
was lower by 4.7%, but OS was similar in all
treatment groups [19]. Finally, the EORTC H9U
study investigated three different chemo regimens all followed by consolidative 30–40 Gy
IFRT. The results showed that ABVD × 4 cycles
and BEACOPP × 4 cycles were not inferior to
ABVD × 6 cycles with 5-year EFS of 86%, 89%,
and 90%, thus leading to the conclusion that
ABVD × 4 cycles followed by IFRT yields high
175
disease control in early unfavorable HL [20].
These large trials of the GHSG and the EORTC
have established combined-­
modality therapy
with reduced-field RT as the treatment of choice
for patients with stage I–II disease.
Recently, trials utilizing results of interim
PET scans that were performed after two or three
cycles of ABVD to identify possible patients who
may be treated with chemotherapy alone have
been reported [21–23]. In the UK RAPID trial,
researchers tested a chemotherapy-alone treatment program for patients with favorable stage
I–II HL who had a negative PET, defined strictly
as Deauville 1–2 only, after three cycles of
ABVD. They found that ABVD × 3 cycles was
inferior to combined-modality therapy in a per-­
protocol analysis in which randomized groups
were analyzed as treated; progression-free survival was significantly better for patients who
received consolidative RT (HR 2.36 in favor of
IFRT, p = 0.02) [23]. Most recently, in data presented at the International Symposium on
Hodgkin Lymphoma 2018 meeting (ISHL11) in
Cologne, Germany, additional analysis of the UK
RAPID study found that as maximum tumor
diameter (MTD) increased, so did the risk of
relapse, specifically in patients who did not
receive RT. For patients with an MTD < 5 cm,
5-year event-free survival was 93.6% where as it
was 79.3% in patients with an MTD ≥ 5 cm (HR
1.23 [95% CI: 1.01–1.48], p = 0.04) [24].
Similarly, a chemotherapy-alone approach has
been proven inferior by the EORTC H10 trials for
patients with favorable and unfavorable stage I–II
HL. In the EORTC H10F and H10U trials, the
ABVD-alone arms for patients who were PETnegative (Deauville <3) after ABVD × 2 cycles
were terminated early due to an excess number of
events when radiation therapy was not incorporated into the therapy even though RT omission
was compensated for by an intensification in the
number of cycles of ABVD [22]. In the final analysis of the favorable subset of H10, ABVD with
INRT resulted in a 5-year PFS of 99.0%, while
ABVD alone resulted in a 5-year PFS of only
87.1%. Similarly, in the unfavorable subset, noninferiority also could not be demonstrated with
chemotherapy alone as the 5-year PFS was 92.1%
in the combined-modality arm and 89.6% in the
J. Yahalom et al.
176
chemotherapy-alone arm. Thus, H10 concluded
that combined modality with ABVD and INRT
remained the standard of care for patients
with either favorable or unfavorable stage I–II
HL [25].
Most recently, the GHSG presented the results
of HD16 at the ISHL11 and the American Society
of Hematology (ASH) 2018 meetings. Patients
with early-stage favorable HL were randomized
to either a standard arm of ABVD × 2 cycles followed by 20 Gy IFRT versus an experimental
arm of no further therapy if they were PET-­
negative, defined as Deauville 1–2, at the end of
two cycles of ABVD. The initial analysis, which
included Deauville 1–3 patients, showed the
omission of RT resulted in inferior outcomes for
these favorable-risk patients with a 5-year estimated PFS of 93.4% in the combined-modality
arm and 86.1% in the chemotherapy-alone arm.
The difference of −7.3% [95% CI: −13.0%,
−1.6%] could not exclude the prespecified non-­
inferiority margin of 3.01%, and thus, combined
modality with ABVD × 2 cycles followed by
20 Gy of radiation remains the standard of care
for patients with early-stage favorable HL [26].
Finally, the UK RATHL trial, though advertised as a trial for advanced-stage HL, actually
was comprised of approximately 50% of patients
with stage II disease [27]. This included patients
with B symptoms, large mediastinal adenopathy,
or >2 sites of disease. Patients with a negative
interim PET (Deauville <4) were treated with
ABVD or ABVD/AVD chemotherapy alone,
without RT. The 3-year PFS was a respectable
90.0%, and the authors conclude that this is
acceptable. However, patients on the H10U trial
who were treated with just four cycles of ABVD
followed by consolidative RT had a 3-year PFS
of 95% (Table 9.1).
Thus, we have learned from GHSG HD8 that
reducing the irradiated volume from the extended-­
field RT that was used in the era before adequate
systemic therapy to involved-field RT does not
result in inferior outcomes. We have also learned
from GHSG HD10 and HD11 that combined
modality with reduced dose and reduced-volume
RT after systemic therapy results in excellent outcomes for patients with early-stage HL. Finally,
we may conclude from the UK RAPID, EORTC
H10, and GHSG HD11 trials that the omission of
RT results in inferior outcomes and combined
modality remains the standard of care. This conclusion has been further bolstered by a recent
systematic review, in which combined-modality
treatment was found to improve tumor control
and overall survival in patients with early-stage
Table 9.1 Summary of trials for stage I–II Hodgkin lymphoma in the PET era (Courtesy of Dr. Richard Hoppe,
Stanford University, United States of America)
Study
NCIC CTG HD.6 [28]
Definition of
PET negative
CT CR/cru
RAPID [23] (per
protocol)
EORTC/GELA/FIL
H10F [25]
EORTC/GELA/FIL
H10U [25]
GHSG HD16 [26]
D<3
Israeli [29]
D<4
CALGB/Alliance
50604 [30]
RATHL [27]
D<4
Total chemo
CR/cru ABVD × 4
PR ABVD × 6 (5)
ABVD × 3
ABVD × 3 + RT
ABVD × 4
ABVD × 3 + RT
ABVD × 6
ABVD × 4 + RT
ABVD × 2 + RT
ABVD × 2
ABVD × 2–4 + RT
ABVD × 4–6
ABVD × 4
D<4
A(B)VD × 6
D<3
D<3
D<3
PFS (%)
(years)
95 (5)
81.0 (5)
97.1 (3)
90.8 (3)
87.1 (5)
99.0 (5)
89.6 (5)
92.1 (5)
93.4 (5)
86.1 (5)
98.5 (5)
88.6 (5)
92.0 (3)
90.0 (3)
PFS
diff
OS
94 (12)
6.3
97.1 (3)
11.9
100 (5)
99.6 (5)
98.3 (5)
96.7 (5)
98.1 (5)
98.4 (5)
2.5
7.3
Notes
Excludes B sx,
bulk
Excludes B sx,
bulk
EORTC
favorable
EORTC
unfavorable
GHSG
favorable
9.9
Non-­
randomized
B sx, bulk, >
sites
9
Principles of Radiation Therapy for Hodgkin Lymphoma
Hodgkin lymphoma [31, 32]. We acknowledge
there may be select early-stage HL patients for
whom a chemotherapy-alone approach may be
preferred. A commonly cited example is a young
woman who would likely receive a large volume
of radiation to breast tissue due to her anatomy or
localization of disease in the mediastinum and
axillae. It is our recommendation that these
patients are discussed in a multidisciplinary conference prior to the start of treatment and that
patients are made aware of all possible treatment
options so that their preferences may be
considered.
9.3.3
Stage III–IV HL
Although the role of consolidative RT after
induction chemotherapy in stages III–IV remains
controversial, RT is often added in patients who
present with bulky disease or who do not have a
clear complete remission after chemotherapy
[33]. The results of prospective studies testing the
concept have been conflicting. A meta-analysis
of several randomized studies demonstrated that
the addition of radiotherapy to chemotherapy
reduces the rate of relapse but did not show survival benefit for combined modality compared to
chemotherapy alone [34]. Unfortunately, nearly
all studies that addressed the question of adding
RT in stage III–IV disease were conducted in the
pre-PET era. With interim PET imaging, it is possible that a more selective use of RT would prove
its benefit.
For historical context, we will briefly discuss
three main pre-PET era studies. The EORTC
20884 trial was a randomized study that evaluated the role of IFRT in patients with stage III–IV
Hodgkin disease who obtained a CR after MOPP/
ABV chemotherapy [35]. Patients received six or
eight cycles of MOPP/ABV (number of cycles
depended upon the response). Patients who did
not achieve a CR (40%) based upon CT imaging
only were not randomized but were all assigned
to receive IFRT. Among the 333 randomized
patients, the 5-year overall survival rates were
91% (no RT). Among the partial responders after
six cycles of MOPP/ABV, the addition of IFRT
177
yielded overall survival and event-free survival
rates that were similar to those obtained among
patients who achieved a CR to chemotherapy.
This suggests a key role for consolidative RT in
stages III–IV when patients fail to achieve a complete response to chemotherapy. Unfortunately,
MOPP/ABV is toxic and has been abandoned for
use in North America. A more modern randomized study evaluated the role of consolidation RT
after CR to chemotherapy used ABVD × 6 cycles,
which is the most common regimen currently
used for advanced-stage HL. This trial was conducted at the Tata Medical Center in India. It
included patients of all stages, but nearly half
were stages III–IV. A subgroup analysis of these
patients showed a statistically significant
improvement of both 8-year event-free survival
(EFS) and 8-year overall survival with added RT
compared to ABVD alone (EFS 78 vs. 59%;
p < 0.03 and OS 100 vs. 80%; p < 0.006) [36].
Finally, a secondary analysis of the UKLG LY09
study evaluated the effect of consolidation RT
following different chemotherapy regimens in
advanced-stage patients. Although more patients
with bulky disease and partial response were in
the RT group, PFS and overall survival were significantly better for 43% of the patients who
received RT in this study. Subgroup and multivariate analysis confirmed this benefit from additional RT [37].
The first study to incorporate PET imaging in
an attempt to define the more selective use of RT
was the GHSG HD15 trial. In this trial, patients
with advanced disease were treated with different
schedules
of
BEACOPP
chemotherapy.
Following completion of chemotherapy, patients
with residual disease greater than 2.5 cm underwent PET imaging. If the PET scan was negative,
patients received no further therapy. If the PET
scan was positive, the patients received 30 Gy of
consolidative RT. Although the group with a positive PET scan had a worse PFS than the PET-­
negative group (86.2% vs. 92.6%), the results in
the PET-positive group were actually quite good
for this subset of poor-prognosis patients, supporting the use of RT for patients in PR by PET
following completion of chemotherapy [38].
Another recent trial evaluated the role of RT
J. Yahalom et al.
178
among stage IIB–IVB patients who had interim
and end-of-treatment PET-negative disease [39].
The GITIL/FIL HD 0607 trial randomized
patients who had nodal disease >5 cm to receive
no further treatment or consolidative RT to the
initially bulky sites following ABVD × 6 cycles.
With a median follow-up of 3.6 years, there was
no significant difference in PFS (93% vs. 97% at
3 years, p < 0.3) or OS (99% vs. 100%, p < 0.08).
The RT question was not addressed for patients
who failed to achieve a complete metabolic
response following the completion of chemotherapy. Finally, the US Intergroup trial S0816 treated
patients with chemotherapy alone, including chemotherapy escalation for patients who had an
interim-positive PET [40]. Among patients who
had an end-of-treatment positive PET, the 2-year
PFS was only 30.6%. Although no RT was utilized for these patients, an analysis was completed assuming patients who met the GHSG
HD15 criteria were irradiated [41]. Assuming a
modest 50% local control for RT, this would have
boosted the likely PFS from 30.6% to 42.8%.
Assuming a more likely 80% local control for
RT, this would have boosted the 2-year PFS to
50.2%.
In summary, the data from the EORTC 20884
and GITIL/FIL HD 0607 suggest a limited role
for RT among patients who achieve a complete
response to chemotherapy. In contrast, the
EORTC 20884, GHSG HD15, and special analysis of the US Intergroup S0816 trial all suggest
that patients who fail to achieve a CR to chemotherapy are very likely to be benefited by the
incorporation of RT. Some patients may benefit
simply from consolidative RT at the conclusion
of chemotherapy, while others may benefit from
its inclusion in an overall salvage treatment
program.
9.3.4
T in Salvage Programs
R
for Refractory and Relapsed HL
High-dose therapy supported by autologous stem
cell transplantation (ASCT) has become a standard salvage treatment for patients with HL who
relapse or remain refractory to primary therapy.
Many of these patients have not received prior RT
or have relapsed at sites outside the original radiation field. These patients could benefit from
integrating RT into the salvage regimen.
Poen and colleagues from Stanford analyzed
the efficacy and toxicity of adding cytoreductive
or consolidative RT to 24 of 100 patients receiving high-dose therapy [42]. When involved sites
were irradiated in conjunction with transplantation, no in-field failures occurred. While only a
trend in favor of IFRT could be shown for the
entire group of transplanted patients, analysis
restricted to patients who had no prior RT or
those with relapse stages I–III demonstrated significant improvement in freedom from relapse.
Fatal toxicity in this series was not influenced
significantly by IFRT. Similar improvements in
outcomes by the addition of RT have been demonstrated in multiple other series including
­studies from the University of Rochester [43] and
the University of Torino [44]. At MSKCC, a program that integrated RT into the high-dose regimen for salvage therapy was developed and
included accelerated hyperfractionated irradiation (twice daily fractions of 1.8 Gy each) to start
after the completion of reinduction chemotherapy and stem cell collection and prior to the highdose chemotherapy and stem cell transplantation
[45–47]. Patients who had not been previously
irradiated received involved-field RT (18 Gy in
5 days) to sites of initially bulky (>5 cm) disease
and/or residual clinical abnormalities, followed
by total lymphoid irradiation (TLI) of 18 Gy
(1.8 Gy per fraction, bid.) during an additional
5 days. Patients who had prior RT received only
involved-­field RT (when feasible) to a maximal
dose of 36 Gy. A recent report detailed the outcomes of 186 patients treated from 1985 to 2008.
The 10-year OS and EFS were 56% [48]. The
authors concluded that this was a safe and effective salvage strategy. A report on the quality of
life and treatment-related complications of this
program disclosed only a small number of late
complications [49].
ILROG has published consensus guidelines
regarding best practice for inclusion of RT in salvage treatment programs for Hodgkin lymphoma
[50]. This report details the patient variables that
9
Principles of Radiation Therapy for Hodgkin Lymphoma
affect selection of salvage treatment, including
intensity of prior therapy, extent of relapse,
whether disease is chemorefractory, and how
radiation can best be incorporated into effective
salvage therapy.
9.4
Mantle
Radiation Fields
and Volumes: Principles
and Design
In the past, radiation-field design attempted to
include multiple involved and uninvolved lymph
node sites. The large fields known as mantle,
inverted Y, and TLI were synonymous with the
radiation treatment of HL. These fields are no
longer in use.
IFRT encompasses a significantly smaller volume and was incorporated into many clinical trials of the past two decades. Extending this
concept further, even more limited radiation volumes termed involved-node radiation therapy
(INRT) and involved-site radiation therapy
(ISRT) have been introduced into combined-­
modality programs and endorsed by guideline
groups as the new standard RT volumes for HL
[9, 10]. Even when radiation is used as primary
management for LPHL, the treatment volumes
should be limited to the involved site or to the
involved sites and immediately adjacent the
lymph nodes.
The terminologies that define radiation volumes may be confusing and create difficulties in
comparing treatment programs. However, general definitions and guidelines are now available
and should be followed [9]. The following are
definitions of types of radiation fields and volumes that have been used in HL.
9.4.1
179
Extended-Field Radiation
Therapy
This field includes the involved lymph node
group plus the adjacent clinically uninvolved
region(s). For extranodal disease, it includes the
involved organ plus the clinically uninvolved
lymph node region. It was common during the
Paraaortic
Pelvic
Fig. 9.1 Illustration of extended RT fields used in the
past
era of treatment with RT alone to treat large
fields encompassing multiple lymph node
regions, both involved and uninvolved. The field
design that includes all of the supradiaphragmatic lymph node regions was referred to as the
mantle field. The field that includes all lymph
node sites below the diaphragm (with or without
the spleen and called after its shape) is the
inverted Y. When all the major lymph node
regions above and below the diaphragm were
irradiated, this was referred to as total lymphoid
irradiation (Fig. 9.1). If the pelvic nodes were
not included, this was referred to as subtotal
nodal irradiation. Extended fields are rarely
used in modern treatment of HL.
9.4.2
Involved-Field Radiation
Therapy
These fields are limited to the clinically involved
lymph node regions [51]. It was influenced by
lymphoid regions that were defined in the Ann
Arbor staging system for Hodgkin’s disease
[52]. For extranodal sites, the field includes the
organ alone (if no evidence for lymph node
J. Yahalom et al.
180
involvement). IFRT was commonly employed
in clinical trials during the past two decades, but
fields have now become even smaller as 3D
cross-sectional imaging has become widely
available. The new volumes based on CT-based
simulation are involved-node radiation therapy
(INRT) and involved-site radiation therapy
(ISRT).
9.4.3
Involved-Site Radiation
Therapy (ISRT): The New
Standard Volume for HL
The International Lymphoma Radiation
Oncology Group (ILROG) now recommends
the use of ISRT to treat HL [9]. ISRT has
already been adopted as the standard volume
by several organizations including the NCCN
[2]. In the majority of cases, assuming the
same clinical presentation and response, ISRT
is smaller than IFRT and more precise as treatment volumes are determined by modern
cross-sectional imaging such as CT and
PET-CT rather than by standard bony landmarks of the involved location as seen on 2D
imaging. The concept of ISRT was developed
as an extension of the INRT concept that was
conceived earlier [10]. In comparison to INRT,
ISRT allows for more flexibility and use of
clinical judgment when the strict criteria for
INRT pre-chemotherapy imaging cannot be
met. Indeed, in the majority of practices, preresection or pre-chemotherapy precise imaging
is not available in the radiation treatment position. ISRT accounts for this deficiency. INRT
is fundamentally a more optimal case of ISRT
when accurate pre-chemotherapy imaging
allows for tighter margins around the original
volumes. Finally, unlike IFRT, which uses predetermined anatomical regional “borders”
determined by bony landmarks that are easy to
visualize during conventional 2D simulation,
which has now been replaced by CT or PET/
CT simulation, ISRT and INRT incorporate the
current concepts of volume determination as
outlined in the ICRU Report 83 [53]. The modern RT treatment volumes are based on defining a gross tumor volume (GTV), a clinical
target volume (CTV), and a planning target
volume (PTV). The PTV is then used to define
beam coverage.
9.4.3.1 I SRT When RT Is the Primary
Treatment
RT as single modality in HL is relevant for
stage I–II lymphocyte-predominant Hodgkin
lymphoma (LPHL). It may also be relevant in
selected cases of early-stage classic HL in
patients who are not candidates for primary
chemotherapy due to serious comorbidities. In
most clinical situations that require RT as the
primary modality, the GTV should be readily
visualized during simulation. In this situation,
the clinical target volume (CTV) should be
more generous since microscopic or subclinical
disease is more likely to be present without
chemotherapy.
9.4.3.2 I SRT When RT Is Part
of Combined-Modality
Treatment
RT is often part of the treatment program for
early-stage classic HL following adequate systemic chemotherapy. RT improves freedom
from treatment failure and progression-free survival even in patients with a negative interim
PET [22, 23, 26,] and allows for a reduced number of chemotherapy cycles [18]. In a recent
systematic review, combined-modality treatment was found to improve tumor control and
overall survival in patients with early-stage
Hodgkin lymphoma [29]. In select patients with
advanced-stage disease, localized RT may be
used for residual sites of lymphoma after full
course of chemotherapy [39]. The GTV may be
markedly affected by prior systemic chemotherapy, and it is therefore particularly important to
review the pre-­chemotherapy imaging and to
define the pre-­
chemotherapy volume on the
9
Principles of Radiation Therapy for Hodgkin Lymphoma
simulation CT study as “pre-chemotherapy
GTV” as well as the post-chemotherapy remaining CT and/or PET abnormality as “post-chemotherapy” GTV.
181
INRT was originally developed and implemented by the EORTC to replace IFRT in prospective randomized studies (EORTC/GELA/
IIL H10). It mandated accurate PET/CT information prior to chemotherapy and in a position similar the subsequent post-chemotherapy
radiation therapy treatment position. The
INRT technique reduces the treated volume to
a minimum, but in order to be safe, optimal
imaging both before and after chemotherapy is
needed [9, 38]. INRT represents a special case
of ISRT, where pre-chemotherapy imaging is
ideal for post-chemotherapy treatment plan-
ning (Fig. 9.2). PET/CT up front for staging
purposes is mandatory as it has been demonstrated that PET/CT is the most accurate imaging method for determining disease extent in
HL [39]. In order to enable image fusion of the
pre-­chemotherapy and the post-chemotherapy
planning images, the pre-chemotherapy PET/
CT scan should be acquired with the patient in
the treatment position and using the same
breathing instructions that will be used later
for RT. Ideally, the patient should be scanned
on a flat couch top, with the use of appropriate
immobilization devices and using markers at
skin positions which are visible in the imaging. During or following the completion of
chemotherapy, a response assessment using
PET/CT or contrast-enhanced CT should be
performed. A planning CT scan is acquired
with the patient in the same position as in the
pre-chemotherapy CT scan. This highly conformal treatment technique has been shown to
be safe, provided strict adherence to the principles above is maintained [54–56].
Fig. 9.2 Involved-node radiation therapy. Single lymph
node in the left lower neck prior to chemotherapy (left) and
following chemotherapy (right). The border of the field
encompasses the original volume of the node and not of the
whole unilateral neck (as in IFRT approach) (Courtesy of
Dr. Theodore Girinsky, Institute Goustave-­Roussy, France)
9.4.4
Involved-Node Radiation
Therapy (INRT): A Special Case
of ISRT
J. Yahalom et al.
182
9.4.5
Volume Definitions
for Planning ISRT and INRT
These principles apply regardless if RT is used
as primary treatment or as part of combined
modality and are relevant to both involved-site
radiation therapy (ISRT) and involved-node
radiation therapy (INRT). The only difference
between ISRT and INRT is the quality and accuracy of the pre-­chemotherapy imaging, which
determines the margins needed to allow for
uncertainties in the contouring of the clinical target volume (CTV).
9.4.5.1 Volume of Interest Acquisition
Planning RT for lymphoma is based on obtaining
a three-dimensional (3D) simulation study using
either a CT simulator, a PET/CT simulator, or an
MRI simulator. If PET and/or CT information
has been obtained separately or prior to simulation, it is possible to transfer the data either manually or electronically into the simulation CT
data. Ideally, any imaging studies that provide
planning information should be obtained in the
treatment position and using the planned immobilization devices.
9.4.5.2 Determination of Gross Tumor
Volume (GTV)
Pre-chemotherapy (or Presurgery) GTV
Any abnormalities on imaging studies obtained
prior to any intervention that might have affected
lymphoma volume should be outlined on the simulation study, as these volumes should (in most
situations) be included in the CTV.
No Chemotherapy or Post-chemotherapy
GTV
The primary imaging of untreated lesions or
post-chemotherapy residual GTV should be outlined on the simulation study and is always part
of the CTV.
9.4.5.3 Determination of Clinical
Target Volume (CTV)
CTV encompasses in principle the original (prior
to any intervention) GTV. Yet, normal structures
such as the large vessels, lungs, kidneys, and
muscles that were clearly uninvolved should be
excluded from the CTV based on clinical judgment. In outlining the CTV, the following points
should be considered:
(a) Quality and accuracy of imaging and transfer
of volumes to simulation images.
(b) Concerns of changes in volume since imaging.
(c) Patterns of spread.
(d) Potential subclinical involvement.
(e) Adjacent organs constraints.
If separate nodal volumes are involved, they
can potentially be encompassed in the same
CTV. However, if the involved nodes are >5 cm
apart, they can be treated with separate volumes
using the CTV-to-PTV expansion guidelines as
outlined further.
9.4.5.4 Determination of Internal
Target Volume (ITV)
ITV is defined in the ICRU Report 62 [54] as the
CTV plus a margin that accounts for uncertainties
in size, shape, and position of the CTV within the
patient. The ITV is mostly relevant when the target is moving with respiration, most commonly in
the chest and upper abdomen. The optimal way to
manage respiratory motion is to use 4D-CT simulation to understand target movement and to generate accurate ITV margins or to use breath-hold
techniques. Alternatively, the ITV may be determined by fluoroscopy or estimated by an experienced clinician. In the chest or upper abdomen,
margins of 1.5–2 cm in the superior-­
inferior
direction may be necessary. In sites such as the
neck, which are well immobilized and unlikely to
change shape or position during or in between
treatments, outlining the ITV is not required.
9.4.5.5 Determination of Planning
Target Volume (PTV)
PTV is the volume that considers the CTV (or
ITV, when relevant) and also accounts for setup
uncertainties in patient positioning and alignment
of the beams during treatment planning and
through all treatment sessions. The practice of
determining the PTV varies across institutions.
9
Principles of Radiation Therapy for Hodgkin Lymphoma
The clinician and/or treatment planner adds the
PTV and applies standard margins that depend on
estimated setup variations that are a function of
immobilization device, body site, and patient
cooperation. While standard patient setup has
historically been done based on skin marks and
weekly portal films, daily image-guided RT
(IGRT) has allowed for reduction in PTV margins. The smaller margins are a function of more
certainty with setup, which can ultimately help
reduce the radiation dose to the normal structures. IGRT can include daily orthogonal KV
images or cone beam CT scan (KV or MV). In
general, PTV expansions can range from 0.3 to
1.0 cm depending on the location and use of
IGRT.
9.4.6
Determination of Organs
at Risk (OAR)
The OARs are normal structures that, if irradiated, could result in significant morbidity including acute toxicity, such as pneumonitis and
esophagitis, and late toxicity, such as hypothyroidism, cardiac toxicity, and second cancers.
OARs may influence treatment planning or the
prescribed dose. They should be outlined on the
simulation study. Dose-volume histograms
(DVH) and normal tissue complication probability (NTCP) should be calculated by the planner
and the plan vetted by the clinician in consideration of this information. Of note, the general
principle with regard to OARs in HL should
always be ALARA—as low as reasonably
achievable—and should depend on the disease
distribution, planned treatment volume, and total
dose.
9.4.6.1 Lung
A major concern for patients with HL and mediastinal disease who receive RT is pneumonitis with
grade 3 pneumonitis rates as high as 7% reported
by the MD Anderson Cancer Center (MDACC)
following IMRT. This is especially true if given as
part of second-line therapy and transplant. The
MDACC group evaluated factors predictive of
grade 1–3 pneumonitis (14% overall) and found
183
the risk of radiation pneumonitis increases with
mean lung dose >13.5 Gy, V20 > 30%, V15 > 35%,
V10 > 40%, and V5 > 55%. Of note, the strongest
predictor was V5 with V5 > 55% associated with a
risk of pneumonitis of almost 35% [57].
9.4.6.2 Heart
Multiple studies have explored the relationship of
heart dose and cardiac toxicity and death among
HL survivors. Cardiac toxicity may be due to pericarditis, arrhythmia, coronary artery disease, valvular disease, and cardiomyopathy/congestive
heart failure. Dosimetric factors have been identified that have been associated with increased risk
of cardiac toxicity. Van Nimwegen et al. reported
on a cohort of HL survivors from the Netherlands
and found a mean heart dose correlated well with
coronary heart disease, demonstrating an excess
relative risk of 7.4% per Gy mean heart dose [58].
A statistically significant increased risk of coronary
heart disease was demonstrated among patients
getting a mean heart dose as low as 5–14 Gy (RR
2.31) compared with a mean heart dose of 0 Gy.
This risk was even higher for mean heart dose of
15 Gy or higher (RR 2.83 for 15–19 Gy, 2.9 for
20–24 Gy, and 3.35 for 25–34 Gy).
A recent analysis of 24,214 5-year survivors
of childhood cancer in the Childhood Cancer
Survivor Study provided substantial insights into
the relationships between radiation and risk of
long-term cardiac disease. Mean heart doses
>10 Gy were associated with increasing cardiac
disease risk in a dose-response manner. Volumes
of the heart receiving radiation also were correlated with cardiac risk; children receiving a V5 of
>50% had a 1.6-fold increased risk of late cardiac
disease. Those receiving at least 20 Gy to any
part of the heart also were at increased risk.
Current recommendations are to keep the mean
heart dose as low as possible with stricter goals to
try and keep the mean heart dose <15 Gy in adults
and <10 Gy in pediatrics whenever possible.
Rarely, should mean heart doses greater than
20 Gy be used, unless patients are being treated
definitively in the salvage setting [59].
While mean heart dose may be appropriate for
radiation evaluation for IFRT, it is unclear whether
it is as important when more conformal techniques
J. Yahalom et al.
184
are being used, which can redistribute the dose
[60]. Recent studies have demonstrated radiation
dose relationships with specific cardiac substructures. For example, Van Niwmegen demonstrated
a relationship between heart failure and mean dose
to the left ventricle [56]. Among patients treated
with anthracyclines, the 25-year cumulative risk of
heart failure was 11.2% for mean LV dose <15 Gy,
15.9% for 16–20 Gy, and 32.9% for greater than or
equal to 21 Gy. Another study by Cutter et al. demonstrated 30-year cumulative risks of valvular
heart disease of 3%, 6.4%, 9.3%, and 12.4% for
mean valvular dose of <30, 31–35, 36–40, and
>40 Gy [61]. Based on this data, we would recommend keeping the mean valve dose <30 Gy, mean
left ventricle dose <15 Gy, and ideally <5 Gy.
9.4.6.3 Thyroid
While hypothyroidism is a common late toxicity, it can be easily managed with thyroid supplementation medication. Cella et al.
demonstrated a dose volume effect with 11.5%
of patients with hypothyroidism with a
V30 < 63% vs. 71% for patients with V30 greater
than or equal to 63% [62].
9.4.6.4 Second Cancers
The primary cause of death among long-term survivors of HL is second cancers. The most common
among survivors are breast cancer, thyroid cancer,
sarcomas (bone and soft tissue), and lung cancer.
Risk modeling has identified linear dose risks for
all these cancers, except for thyroid cancer.
Breast cancer is the most concerning second
cancer among female survivors receiving radiation. While smaller treatment fields have greatly
reduced the risk of second breast cancers, consideration of breast dose is important in minimizing
the risk for all patients. Travis et al. demonstrated
that radiation doses >4 Gy were associated with
increased risk of secondary breast cancer with
increasing dose further increasing the risk. In
fact, the relative risk was 1.8 and 4.1 for breast
dose of 4–6.9 and 7–23.1 Gy, respectively [63].
Therefore, it is important to keep the mean breast
dose as low as possible and try to minimize the
breast V4 to as low as possible.
Lung cancer is an aggressive second cancer
that will often result in death for a HL survivor.
Fortunately, the risk can greatly be mitigated for
nonsmokers. However, among smokers, this risk
is increased significantly with the addition of lung
irradiation. Travis et al. demonstrated among HL
survivors that lung dose of >5 Gy had a r­elative
risk of 5.9 compared with lung dose <5 Gy for
developing a second lung cancer [64]. Similar to
concerns of pneumonitis, lung V5 should be evaluated with attempts to keep as low as possible.
Secondary sarcomas have also been found to
increase with higher radiation doses to the body.
Tukenova et al. demonstrated increased risk
(12.5) for dying from a secondary sarcoma when
the calculated integral dose was >150 J [65].
Thyroid cancers do not have a linear dose-risk
relationship with radiation. Bhatti et al. demonstrated a relative risk increase in secondary thyroid cancers of 8.5 with doses of 5–10 Gy, 10.6
for 10–15 Gy, 13.8 with 15–20 Gy, and 14.6 for
20–25 Gy, with relative risks then declining with
doses >25 Gy [66].
9.4.7
Consolidation Volume
Radiation Therapy (CVRT)
As systemic therapies continue to improve, further reductions in RT volumes may be possible.
Currently, there are ongoing studies looking at
modifications in systemic therapy to include
brentuximab vedotin (BV) and/or checkpoint
inhibitors, such as pembrolizumab and nivolumab
[67, 68]. One recent reported study from
Memorial Sloan Kettering Cancer Center
(MSKCC) demonstrated a favorable response to
BV-AVD × 4 cycles followed by ISRT [69].
Consolidation volume radiation therapy (CVRT),
which treats only residual CT abnormalities in
patients who achieve a CR by PET, is currently
being tested after BV-AVD × 4 cycles [70].
9.5
Dose Considerations
and Recommendations
Although doses in the range of 40–44 Gy were at
one time recommended for the definitive treatment of HL, these recommendations have been
modified over time, both in the context of
9
Principles of Radiation Therapy for Hodgkin Lymphoma
combined-­
modality therapy for cHL and the
treatment of patients with LPHL. The radiation
dose is typically delivered in 1.8–2.0 Gy fractions. If significant portions of lung or heart are
included, the dose per fraction can be reduced to
1.5 Gy. The available data indicate that the choice
of fractionation is not critical for tumor control
and that a schedule with minimal risk of damage
to normal structures should be selected.
The GHSG evaluated dose in patients with
stage IA–IIB disease without risk factors in a randomized trial of 40 Gy extended-field radiation
alone vs. 30 Gy extended-field radiation with a
boost of 10 Gy to the involved site of disease [71,
72]. There was no significant difference in outcome between the two arms of the study indicating
that 30 Gy is sufficient for clinically uninvolved
areas when RT is used alone. The optimum dose
for clinically involved sites of disease with RT
alone has not been tested in a randomized trial.
More relevant to current practice is the determination of the adequate radiation dose after treatment with chemotherapy. In many early studies,
radiation doses were kept at approximately 40 Gy
even after achieving a CR to chemotherapy; others
reduced the dose in advanced disease when combined with five cycles of chemotherapy to 20–24 Gy
with excellent overall results [73]. Studies of combined modality in advanced stage also used reduced
doses of RT for patients who achieved a CR to chemotherapy and higher doses (approximately 30 Gy)
for patients in PR. The pediatric groups addressing
the concern of radiation effects on skeletal and
muscular development also effectively reduced the
dose of RT after combination chemotherapy to
21–24 Gy [74]. However, recent reports have demonstrated that >90% of all relapses occur in field,
suggesting higher doses may be appropriate for
pediatrics in certain cases [75].
Several recent studies addressed the adequacy
of low-dose IFRT following chemotherapy. A
study conducted by the EORTC/GELA [76] randomized patients with favorable early-stage HL to
36, 20, or no IFRT after achieving a CR to six
cycles of EBVP. Because an excessive number of
relapses occurred in the no-RT arm, this arm was
closed early. There was no difference in EFS at
4 years between patients receiving IFRT 36 Gy
(87%) vs. 20 Gy (84%). A GHSG randomized
185
study (HD 10) addressed the radiation dose question after short-course chemotherapy [18]. Patients
with favorable stages I–II were randomized to
receive either four or only two cycles of ABVD followed by IFRT of 30 or 20 Gy. At a median followup of 7 years, there was no difference in FFTF
among the four arms. FFTF at 5 years was 93.4%
in patients treated with 30 Gy (91.0–95.2%) and
92.9% in those receiving 20 Gy (90.4–94.8%).
These results, taken together with the better tolerability and the lack of inferiority in secondary efficacy endpoints, led to the conclusion that 20 Gy
IFRT, when combined with even only two cycles of
ABVD, is equally effective to 30 Gy IFRT in this
very favorable group of patients [15]. The GHSG
HD11 study targeted patients with unfavorable
early stage and randomized them to either
ABVD × 4 cycles or BEACOPP × 4 cycles; either
program was followed by either 20 or 30 Gy to the
involved field. Five-year FFTF and OS for all
patients were 85% and 94.5%, respectively. There
was
no
difference
in
FFTF
when
BEACOP × 4 cycles was followed by either 30 or
20 Gy, and similar excellent results were obtained
with ABVD × 4 cycles and IFRT of 30 Gy. Patients
who received ABVD × 4 cycles and only 20 Gy
had FFTF that was lower by 4%. OS was similar in
all treatment groups [19]. These results suggest that
30 Gy should remain the standard IFRT dose following ABVD in unfavorable early-stage HL [77].
For patients with early-stage LPHL, no advantage has been shown for doses over 30–35 Gy [15].
For patients with residual lymphoma after
chemotherapy, the residual mass may represent a
more refractory disease, and increasing the dose
to the CTV to 36–40 Gy should be considered.
9.5.1
he Significance of Reducing
T
the Radiation Dose
Recent studies clearly indicate that the risk of secondary solid tumor induction is radiation dose
related. This was carefully analyzed for secondary
breast and lung cancers as well as for other tumors
[63, 64, 78, 79]. While it will take more years of
careful follow-up of patients in randomized studies
to display the full magnitude of risk tapering by
current reduction of radiation volume and dose,
J. Yahalom et al.
186
recent data suggest that this likely to be the case. In
a Duke University study, two groups of patients
with early-stage HL were treated with different
radiation approaches over the same period. One
group received RT alone, given to extended fields
with a median dose of 38 Gy; the second group
received chemotherapy followed by involved-field
low-dose (median of 25 Gy) RT. While 12 patients
developed second tumors in the first group and 8
of them died, no second tumors were detected in
the second group. The median follow-up was 11.7
and 8.1 years, respectively [80]. Similar observations with an even longer follow-up were made by
the Yale group [81]. In a study that used data-based
radiobiological modeling to predict the radiationinduced second cancer risk, lowering the dose
from 35 to 20 Gy and reducing the extended field
to IFRT reduced lung cancer risk and breast cancer
risk by 57% and 77%, respectively [82].
Finally, a study by a French Collaborative
Lymphoma group (GOELAM) randomized
patients with favorable stage I–II HL to receive a
conservative RT dose of 40 Gy to involved sites
and 30 Gy to adjacent site control arm or in the
“experimental arm” to receive only 36 Gy and
24 Gy to the adjacent sites after ABVD × 3 cycles
[83]. Surprisingly, the 10-year incidence of
severe or fatal complications was nil in the experimental arm but reached 15.5% in the control arm
(p < 0.003) and 11.1% in the historical controls
that received the higher dose. The 10-year FFTF
and overall survival rates were similar for the 89
patients in the experimental arm (88.6% and
97.8%, respectively), for the 99 patients in the
conservative arm (92.6% and 95%, respectively),
and for the 202 patients in the historical control
group (91.9% and 92.9%, respectively).
9.5.2
Dose Recommendations
Radiation alone (as primary treatment for LPHL)
using ISRT
• Clinically involved and adjacent uninvolved
nodes: 30–36 Gy.
Radiation alone (as primary treatment for cHL
[uncommon])
• Clinically involved sites: 36 Gy at a
minimum.
• Clinically uninvolved sites: 30 Gy.
Radiation following chemotherapy in a
combined-­modality program
• Patients in CR after chemotherapy: 20–30 Gy.
–– For pediatric or adolescent patients:
15–24 Gy.
–– In some programs of short chemotherapy
for bulky or advanced-stage disease (e.g.,
Stanford V), the recommended RT dose is
30–36 Gy.
• Patients in PR after chemotherapy: 30–40 Gy.
9.6
ew Aspects of Radiation
N
Volume Definition
and Treatment Delivery
The abandonment of large-field irradiation for
most patients with HL permits the use of more
conformal RT volumes and introduction of other
innovative RT techniques. The change in the
lymphoma RT paradigm coincided with substantial improvement in imaging and treatment planning technology that has revolutionized the field
of RT. The integration of fast high-resolution
computerized tomography into the simulation
and planning systems of radiation oncology has
changed how treatment volumes and relationship to normal critical structures are determined
and planned. In the recent past, tumor volume
determinations were made with fluoroscopybased simulators that produced often poor-quality imaging requiring wide “safety margins” that
detracted from accuracy and sparing of critical
organs. Most modern simulators are in fact high-­
resolution CT scanners with software programs
that allow accurate conformal treatment planning and provide detailed information on the
dose volume delivered to normal structures
within the treatment field and the homogeneity
of dose delivered to the target. More recently,
these simulators have been integrated with a
PET scanner that provides additional tumor volume information for consideration during radiation planning.
9
Principles of Radiation Therapy for Hodgkin Lymphoma
9.6.1
New Technologies
Intensity-modulated RT (IMRT), tomotherapy,
and volumetric modulated arc therapy (VMAT)
are advanced systems for photon delivery. These
modalities redistribute the radiation to provide
high-dose conformality of the target area, but with
less conformality in the low-dose region. They
also allow for accurately enveloping the tumor
with either a homogenous radiation dose (“sculpting”) or delivering higher doses to predetermined
areas in the tumor volume (“painting”). In the
treatment of lymphoma, there are several clinical
situations where highly conformal photon techniques provide a benefit. In the mediastinum, a
review article of comparison studies showed that
IMRT compared with conventional 3D-conformal
radiation techniques reduced the mean heart dose
on average by 1.44 Gy, mean esophagus dose by
1.4 Gy, and lung V20 by 11% [84]. Highly conformal photon techniques can also be useful in the
treatment of very large or complicated tumor volumes in the abdomen and head and neck lymphomas. IMRT also allows re-­irradiation of sites prior
to high-dose salvage programs that otherwise will
be prohibited by normal tissue tolerance, particularly of the spinal cord (Figs. 9.3a–d and 9.4a–c).
In general, when using highly conformal photon techniques for mediastinal disease, treatment
planning generally tries to avoid equally spaced
beams or continuous arcs around the patient in an
effort to avoid some of the low-dose bath, especially to the lungs and breasts. MD Anderson
Cancer Center has described both the butterfly
IMRT technique and the rainbow IMRT technique
as ways to optimize the IMRT beam arrangement
in mediastinal lymphoma. In the butterfly technique, three anterior and two posterior beams are
used to reduce excess exposure to heart, lungs,
and spinal cord. In the rainbow technique, one
anterior-posterior (AP) beam and four anterior
obliques at 0°, 20–30°, 40–60°, 300–320°, and
335–345° are used for patients with only anterior
mediastinal disease [85, 86]. Similarly, the
University of Torino evaluated optimized VMAT
plans that include non-­coplanar partial arcs [87,
88]. Early clinical data has begun to emerge from
the use of IMRT for mediastinal lymphoma dem-
187
onstrating similar disease control to 3DCRT treatment [78, 89]. While most have shown little to no
pneumonitis with these t­echniques, MDACC did
report a 15% grade 1–3 pneumonitis risk including 6.9% grade 3 rate [58].
9.6.2
Deep Inspiration Breath Hold
An additional technique that can be used to try
and further minimize heart and lung dose for
mediastinal lymphoma patients and can be used
with 3DCRT, IMRT, or proton therapy is deep
inspiration breath hold (DIBH). DIBH is a simple technique which the patient inhales deeply
and holds this breath during treatment. DIBH
can optimize the internal anatomy by pulling the
heart caudally while allowing the disease to be
irradiated to remain more superiorly by the great
vessels for patients with superior mediastinal
disease. This allows for more cardiac sparing for
these patients. DIBH also immobilizes the disease in the mediastinum, which controls respiratory motion and eliminates the need for an
ITV. Finally, it expands the total lung volume,
which results in an overall decreased dose to the
lungs [90].
Petersen et al. conducted a prospective phase
II study of DIBH among patients with mediastinal lymphoma among 19 patients. In the study,
the mean lung dose was reduced on average by
2 Gy with DIBH and mean heart dose by 1.4 Gy.
Another study by Charpentier et al. reported on
47 patients undergoing DIBH, where the mean
lung dose was reduced by approximately 1.5 Gy
and mean heart dose reduced by 2.5 Gy [91].
While DIBH appears beneficial for disease
located in the superior mediastinum, the benefit
is not as obvious for patients with lower mediastinal disease that extends to the level of the heart
[92]. This was seen in a study by Paumier et al.
who demonstrated a mean heart dose reduction
of 50% for patients with upper mediastinal disease, while it was only 8–9% and not significant
for lower mediastinum. Similarly, the mean lung
dose was reduced by 26% for upper mediastinal
disease, but only 18% reduction for lower mediastinal disease [93] (Fig. 9.5).
188
J. Yahalom et al.
a
b
Fig. 9.3 (a) CT-MR fusion for target localization of HL
involving the mediastinum and right chest wall. CTV clinical treatment volume, PTV planning treatment volume.
(b, c) Treatment plans comparing AP/PA, 3DCRT, and
IMRT. PTV planning treatment volume, AP/PA opposed
anterior and posterior fields, 3DCRT three-dimensional
conformal radiation therapy, IMRT intensity-modulated
radiation therapy. (d) Comparison of lung complication
probability of different plans
9
Principles of Radiation Therapy for Hodgkin Lymphoma
189
c
Plans:
d
AP/PA
3DCRT
IMRT
100
Complication probability
(NTCP) for lungs
Volume (%)
80
60
AP/PA
3DCRT
IMRT
40
20
Lungs
0
0
1,000 2,000 3,000
Dose (cGy)
Fig. 9.3 (continued)
4,000 5,000
19%
16%
3%
190
J. Yahalom et al.
a
b
Fig. 9.4 (a) Use of IMRT for re-irradiation of a patient
relapsing after ABVD and mantle-field irradiation to
36 Gy. (b, c) Treatment planning options for re-­irradiation.
AP/PA opposed anterior and posterior fields, 3DCRT
three-dimensional conformal radiation therapy, IMRT
intensity-modulated radiation therapy
9
Principles of Radiation Therapy for Hodgkin Lymphoma
191
c
Fig. 9.4 (continued)
Fig. 9.5 An example case in which the use of DIBH
increases total lung volume and pulls the heart caudally,
thus decreasing dose to lung and heart without compro-
mising coverage (Courtesy of Lena Specht, MD PhD,
Rigshospitalet, University of Copenhagen, Denmark)
J. Yahalom et al.
192
9.7
Proton Therapy
Another technical advance is the use of particle
therapy (protons). Protons have the advantage of a
more defined depth of penetration than photons,
which eliminates the “exit dose” of photons. Proton
therapy may be helpful in the mediastinum, in select
cases where significant sparing of OARs including
the heart, lungs, and esophagus cannot be achieved
with IMRT [56]. In a review article of several dosimetric studies comparing highly conformal photon
techniques with proton therapy, on average proton
therapy reduced the mean heart dose by 2.24 Gy,
breast by 2.45 Gy, lung by 3.28 Gy, thyroid by
2.09 Gy, and mean body dose by about 40% [76].
Patients with lower mediastinal disease may benefit
more from proton therapy, due to the potential gains
in reducing the radiation dose to the heart as
described in the ILROG guidelines [94]. The
Fig. 9.6 A sample case of an 18-year-old woman with
stage II HL at diagnosis with representative plans using
various treatment modalities including mantle field, IFRT,
ISRT using 3DCRT (ISRT 3D), ISRT using IMRT (ISRT
ILROG guidelines also discuss in detail both the
advantages and disadvantages of proton therapy for
lymphoma and identify parameters that can help
clinicians better select the appropriate modality.
While proton therapy planning and deliver is
more complex than photon-based treatments, multicenter clinical outcomes have demonstrated similar disease control rates with that of IMRT or
3DCRT [95]. Furthermore, risks of pneumonitis
have been extremely low [96], and there can be
improved cardiac sparing [97]. Proton therapy may
also be helpful in other situations including relapsed
and refractory patients that require higher doses of
radiation, when the disease involves the axilla, due
to the ability to spare the breasts with posterior
fields, and in pediatric HL, where the risk of second
cancers is highest. Fig. 9.6 shows representative
colorwash dose distributions for the same patient
across different radiation treatment approaches.
IMRT), and ISRT using proton therapy (ISRT PT)
(Courtesy of Brad Hoppe, MD MPH, University of
Florida, United States of America)
9
Principles of Radiation Therapy for Hodgkin Lymphoma
9.8
ommon Side Effects
C
and Supportive Care During
Radiation Therapy
The side effects of RT depend on the irradiated
volume, dose administered, and technique
employed. They are also influenced by the
extent and type of prior chemotherapy, if any,
and by the patient’s age, habits, and presence of
intercurrent disease. Most of the information
that we use today to estimate risk of RT is
derived from strategies that used radiation
alone, with larger treatment volumes and higher
doses. As noted previously, field sizes have been
reduced and doses decreased, and other technological advances have all drastically reduced the
radiation exposure to the OARs. It is thus misleading to inform patients of risks of RT using
information from RT of the past as this is no
longer practiced.
It is critical to remember that most of the data
of long-term complications associated with RT
and particularly second solid tumors and coronary heart disease were reported from databases
of patients with HL treated more than 25 years
ago. It is also important to note that we have very
limited long-term follow-up data on patients with
HL who were treated with chemotherapy alone.
9.8.1
Common Acute Side Effects
Radiation, in general, may cause fatigue, and
areas of the irradiated skin may develop mild sun
exposure-like dermatitis. The acute side effects
of irradiating the full neck and portions of the
mouth include dryness, change in taste, and
pharyngitis. Patients who are treated to the neck
and mediastinum may also develop mild dysphagia and esophagitis, which is self-limited. With
the doses and techniques of irradiation currently
employed in HL, all of these side effects are usually mild and transient. The main potential side
effects of subdiaphragmatic irradiation are loss
of appetite, nausea, and increased bowel frequency. Again, these reactions are usually mild
and can be minimized with standard antiemetic
medications.
9.8.2
193
Uncommon Early Side Effects
Lhermitte sign: Less than 3% of patients who
have treatment that includes long lengths of the
spinal cord may note an electric shock sensation
radiating down the backs of both legs when the
head is flexed (Lhermitte sign) 6 weeks to
3 months after mantle-field RT. Possibly secondary to transient demyelination of the spinal
cord, Lhermitte sign resolves spontaneously
after a few months and is not associated with
late or permanent spinal cord damage. The risk
is likely increased in the presence of prior neurotoxic chemotherapy such as vincristine or
vinblastine.
Pneumonitis and pericarditis: During the
same period, radiation pneumonitis and/or acute
pericarditis may occur in <3% of patients; these
side effects occur more often in those who have
extensive mediastinal disease. Both inflammatory processes have become rare with modern
radiation techniques.
The consideration and discussion of potential late side effects and complications of both
RT and chemotherapy are of prime importance.
A more complete discussion is detailed in
Chap. 20.
9.8.3
upportive Care During
S
Treatment
It is important to prepare the patient for the
potential side effects of RT, and in addition to
physician-led discussion, many organizations
and cancer centers also provide written patient
information regarding RT for lymphomas. Since
some level of xerostomia may be associated with
RT that involves the upper neck and/or lower
mandible and mouth, attention to dental care is
advised. If dryness is a concern, it is advised to
arrange for a consultation with a dental expert for
overall dental evaluation and consideration of
mouth guards (from scatter) and/or supplemental
fluoride treatment during and after RT.
Soreness of the throat and mild-to-moderate
difficulty of swallowing solid and dry food may
also occur during neck irradiation, with onset at a
J. Yahalom et al.
194
dose of approximately 20 Gy. These side effects
are almost always mild, self-limited, and subside
shortly after completion of RT. Skin care with
hydrating lotion and sunscreen is advised for all
patients undergoing RT. Temporary hair loss is
expected in irradiated areas, and recovery is generally observed after several months.
9.8.4
Follow-Up After Treatment
We normally recommend a first post-RT follow­up visit 6 weeks after the end of treatment and
obtain post-RT baseline blood count, standard
biochemistry tests, as well as TSH levels (if there
was neck irradiation) and lipid profile (if applicable) at that visit. Follow-up imaging studies
normally commence 3 months after completion
of treatment. Patients treated with radiation therapy alone for NLPHL should have a posttreatment PET scan to confirm a complete response.
Other follow-up studies are included in the
NCCN guidelines for HL [2].
References
1. Lebow F (1996) Refining the management of
Hodgkin’s disease. Oncology Times 1996:63
2. Hoppe RT, Advani R, Ai WZ et al (2018) National
comprehensive cancer network guidelines: Hodgkin
lymphoma, version 3.2018
3. Yahalom J (2009) Role of radiation therapy in
Hodgkin’s lymphoma. Cancer J 15:155–160
4. Pusey W (1902) Cases of sarcoma and of Hodgkin’s
disease treated by exposures to x-rays: a preliminary
report. JAMA 38:166–169
5. Senn N (1903) Therapeutical value of roentgen ray in
treatment of pseudoleukemia. N Y Med J 77:665–668
6. Gilbert R (1925) La roentgentherapie de la granulomatose maligne. J Radiol Electrol 9:509–514
7. Peters M (1950) A study in survivals in Hodgkin’s
disease treated radiologically. Am J Roentgenol
63:299–311
8. Kaplan H (1962) The radical radiotherapy of
Hodgkin’s disease. Radiology 78:553–561
9. Specht L, Yahalom J, Illidge T et al (2014) Modern
radiation therapy for Hodgkin lymphoma: field and
dose guidelines from the international lymphoma
radiation oncology group (ILROG). Int J Radiat Oncol
Biol Phys 89(4):854–862. https://doi.org/10.1016/j.
ijrobp.2013.05.005
10. Girinsky T, van der Maazen R, Specht L et al (2006)
Involved-node radiotherapy (INRT) in patients with
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
early Hodgkin lymphoma: concepts and guidelines.
Radiother Oncol 79:270–277
Hodgson DC, Dieckmann K, Terezakis S et al (2015)
Implementation of contemporary radiation therapy
planning concepts for pediatric Hodgkin lymphoma:
guidelines from the international lymphoma radiation
oncology group. Pract Radiat Oncol 5(2):85–92
Brentuximab vedotin and combination chemotherapy in treating children and young adults with stage
IIB or Stage IIIB-IVB Hodgkin lymphoma. (First
posted June 18, 2014, last updated March 12, 2019).
ClinicalTrials.gov Identifier: NCT02166463
Nogova L, Reineke T, Eich HT et al (2005) Extended
field radiotherapy, combined modality treatment or
involved field radiotherapy for patients with stage
IA lymphocyte-predominant Hodgkin’s lymphoma:
a retrospective analysis from the German Hodgkin
study group (GHSG). Ann Oncol 16(10):1683–1687
Wirth A, Yuen K et al (2005) Long-term outcome
after radiotherapy alone for lymphocyte-predominant
Hodgkin lymphoma: a retrospective multicenter study
of the Australasian radiation oncology lymphoma
group. Cancer 104:1221–1229
Chen RC, Chin MS, Ng AK et al (2010) Early-stage,
lymphocyte-predominant Hodgkin’s lymphoma:
patient outcomes from a large, single-institution
series with long follow-up. J Clin Oncol 28:136–141
Schlembach PJ, Wilder RB et al (2002) Radiotherapy
alone for lymphocyte-predominant Hodgkin’s disease. Cancer J 8:377–383
Sasse S, Klimm B, Gorgen H et al (2012) Comparing
long-term toxicity and efficacy of combined modality
treatment including extended- or involved-field radiotherapy in early-stage Hodgkin’s lymphoma. Ann
Oncol 23(11):2953–2959
Engert A, Plutschow A, Eich HT et al (2010)
Reduced treatment intensity in patients with earlystage Hodgkin’s lymphoma. N Engl J Med 363(7):
640–652
Eich HT, Diehl V, Gorgen H et al (2010) Intensified
chemotherapy and dose-reduced involved-field
radiotherapy in patients with early unfavorable
Hodgkin’s lymphoma: final analysis of the German
Hodgkin study group HD11 trial. J Clin Oncol
28(27):4199–4206
Ferme C, Thomas J, Brice P et al (2017) ABVD or
BEACOPPbaseline along with involved-field radiotherapy in early-stage Hodgkin lymphoma with risk
factors: results of the European Organisation for
Research and Treatment of Cancer (EORTC)-Groupe
d’Etude des Lymphomes de l’Adulte(GELA) H9-U
intergroup randomised trial. Eur J Cancer 81:45–55
Evens AM, Kostakoglu L (2014) The role of FDG-­
PET in defining prognosis of Hodgkin lymphoma for
early-stage disease. Blood 124(23):3356–3364
Raemaekers JM, André MP, Federico M et al (2014)
Omitting radiotherapy in early positron emission
tomography-negative stage I/II Hodgkin lymphoma
is associated with an increased risk of early relapse:
clinical results of the preplanned interim analysis of
the randomized EORTC/LYSA/FIL H10 trial. J Clin
9
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
Principles of Radiation Therapy for Hodgkin Lymphoma
Oncol 32(12):1188–1194. https://doi.org/10.1200/
JCO.2013.51.9298
Radford J, Illidge T, Counsell N et al (2015) Results of a
trial of PET-directed therapy for early-stage Hodgkin’s
lymphoma. N Engl J Med 373(17):1598–1607
Illidge T et al (2018) Maximum tumour dimension at
baseline is associated wth event-free survival in PET
negative patients with stage IA/IIA Hodgkin lymphoma in the UK NCRI RAPID trial. International
symposium on Hodgkin lymphoma (ISHL11), 26–29
Oct 2018, Cologne
Andre MPE, Girinsky T, Federico M et al (2017) Early
positron emission tomography response-adapted
treatment in stage I and II Hodgkin lymphoma: final
results of the randomized EORTC/LYSA/FIL H10
trial. J Clin Oncol 35(16):1786–1794
Engert A et al (2018) Early-stage favorable HL:
HD16. International symposium on Hodgkin lymphoma (ISHL11), 26–29 Oct 2018, Cologne
Johnson P, Federico M, Kirkwood A et al (2016)
Adapted treatment guided by interim PET-CT scan
in advanced Hodgkin’s lymphoma. New Engl J Med
374(25):2419–2429
Meyer RM, Gospodarowicz MK, Connors JM et al
(2012) ABVD alone versus radiation-based therapy
in limited-stage Hodgkin’s lymphoma. N Engl J Med
366(5):399–408
Dann EJ, Bairey O, Bar-Shalom R et al (2017)
Modification of initial therapy in early and advanced
Hodgkin lymphoma, based on interim PET/CT is beneficial: a prospective multicenter trial of 355 patients.
Br J Haematol 178(5):709–718
Straus DJ, Jung SH, Pitcher B et al (2018) CALGB
50604: a risk-adapted treatment of nonbulky
early-stage Hodgkin lymphoma based on interim
PET. Blood 132(10):1013–1021
Herbst C, Rehan FA et al (2009) Combined modality
treatment improves tumor control and overall survival
in patients with early stage Hodgkin lymphoma: a
systematic review. Haematologica 95:494
Blank O, von Tresckow B, Monsef I et al (2017)
Chemotherapy alone versus chemotherapy plus radiotherapy for adults with early stage Hodgkin lymphoma. Cochrane Database Syst Rev 4:CD0071010
Loeffler M, Diehl V et al (1997) Dose-response relationship of complementary radiotherapy ­
following
four cycles of combination chemotherapy in intermediate-stage Hodgkin’s disease. J Clin Oncol
15:2275–2287
Skoetz N, Trelle S, Rancea M et al (2013) Effect of
initial treatment strategy on survival of patients with
advanced-stage Hodgkin’s lymphoma: a systematic
review and network meta-analysis. Lancet Oncol
14(10):943–952
Aleman BM, Raemaekers JM et al (2003) Involved-­
field radiotherapy for advanced Hodgkin’s lymphoma.
N Engl J Med 348:2396–2406
Laskar S, Gupta T et al (2004) Consolidation radiation
after complete remission in Hodgkin’s disease following six cycles of doxorubicin, bleomycin, vinblastine,
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
195
and dacarbazine chemotherapy: is there a need? J Clin
Oncol 22:62–68
Johnson PWM, Sydes MR, Hancock BW et al (2010)
Consolidation radiotherapy in patients with advanced
Hodgkin’s lymphoma: survival data from the UKLG
LY09 randomized controlled trial. J Clin Oncol
28:3352–3359
Engert A, Haverkamp H, Kobe C et al (2012)
Reduced-­
intensity chemotherapy and PET-guided
radiotherapy in patients with advanced stage
Hodgkin’s lymphoma (HD15 trial): a randomised,
open-label, phase 3 non-­
inferiority trial. Lancet
379(9828):1791–1799
Gallamini A, Tarella C, Viviani S et al (2018)
Early chemotherapy intensification with escalated
BEACOPP in patients with advanced-stage Hodgkin
lymphoma with a positive interim positron emission
tomography/computed tomography scan after two
ABVD cycles: long-term results of the GITIL/FIL
HD 0607 trial. J Clin Oncol 36(5):454–462
Press OW, Li H, Schoder H et al (2016) US intergroup trial of response-adapted therapy for stage
III to IV Hodgkin lymphoma using early interim
Flurodeoxygluose-positron emission tomography
imaging: southwest oncology group S0816. J Clin
Oncol 34(17):2020–2027
Ha CS et al (2018) Potential impact of consolidation
radiation therapy for advanced Hodgkin lymphoma:
a secondary modeling of SWOG S0816 with receiver
operating characteristic analysis. American Society of
hematology annual meeting, 1–4 Dec San Diego, CA
Poen JC, Hoppe RT et al (1996) High-dose therapy
and autologous bone marrow transplantation for
relapsed/refractory Hodgkin’s disease: the impact of
involved field radiotherapy on patterns of failure and
survival. Int J Radiat Oncol Biol Phys 36:3–12
Biswas T, Culakova E, Friedberg JW et al (2012)
Involved field radiation therapy following high dose
chemotherapy and autologous stem cell transplant
benefits local control and survival in refractory or
recurrent Hodgkin lymphoma. Radiother Oncol
103(3):367–372
Levis M, Piva C, Filippi AR et al (2017) Potential
benefit of involved-field radiotherapy for patients
with relapsed-refractory Hodgkin’s lymphoma with
incomplete response before autologous stem cell
transplantation. Clin Lymphoma Myeloma Leuk
17(1):14–22
Yahalom J, Gulati SC et al (1993) Accelerated hyperfractionated total-lymphoid irradiation, high-dose
chemotherapy, and autologous bone marrow transplantation for refractory and relapsing patients with
Hodgkin’s disease. J Clin Oncol 11:1062–1070
Moskowitz CH, Nimer SD et al (2001) A 2-step comprehensive high-dose chemoradiotherapy second-line
program for relapsed and refractory Hodgkin’s disease: analysis by intent to treat and development of a
prognostic model. Blood 97:617–623
Moskowitz CH, Kewalramani T et al (2004)
Effectiveness of high dose chemoradiotherapy and
J. Yahalom et al.
196
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
autologous stem cell transplantation for patients with
biopsy-proven primary refractory Hodgkin’s disease.
Br J Haematol 124:645–652
Rimner A, Lovie S, Hsu M et al (2017) Accelerated
total lymphoid irradiation-containing salvage regimen for patietns with refractory and relapsed Hodgkin
lymphoma: 20 years of experience. Int J Radiat Oncol
Biol Phys 97(5):1066–1076
Goodman KA, Riedel E et al (2008) Long-term
effects of high dose chemotherapy and radiation for
relapsed and refractory Hodgkin’s lymphoma. J Clin
Oncol 26:5240–5247
Constine LS, Yahalom J, Ng AK et al (2018) The
role of radiation therapy in patients with relapse and
refractory Hodgkin lymphoma: guidelines from the
international lymphoma radiation oncology group. Int
J Radiat Oncol Biol Phys 100(5):1100–1118
Yahalom J, Mauch P (2002) The involved field is back:
issues in delineating the radiation field in Hodgkin’s
disease. Ann Oncol 13(Suppl 1):79–83
Kaplan HS, Rosenberg SA (1966) The treatment of
Hodgkin’s disease. Med Clin North Am 50:1591–1610
ICRU. International Commission on Radiation Units
and Measurements (1999) Prescribing, recording, and
reporting photon therapy. Supplement to ICRU report
50. ICRU report 62
Girinsky T, Specht L, Ghalibafian M et al (2008) The
conundrum of Hodgkin lymphoma nodes: to be or not
to be included in the involved node radiation fields.
The EORTC-GELA lymphoma group guidelines.
Radiother Oncol 88:202–210
Maraldo MV, Aznar MC, Vogelius IR et al (2013)
Involved node radiotherapy: an effective alternative in
early stage Hodgkin lymphoma. Int J Radiat Oncol
Biol Phys 85:1057–1065
Paumier A, Ghalibafian M, Beaudre A et al (2011)
Involved-node radiotherapy and modern radiation
treatment techniques in patients with Hodgkin lymphoma. Int J Radiat Oncol Biol Phys 80:199–205
Pinnix CC, Smith GL, Milgrom S et al (2015) Int J
Radiat Oncol Biol Phys 92(1):175–182
Van Nimwegen FA, Schaapveld M, Cutter DJ et al
(2016) Radiation dose-response relationship for risk
of coronary heart disease in survivors of Hodgkin
lymphoma. J Clin Oncol 34(3):235–243
Bates JE, Howell RM, Liu Q et al (2019) Therapy-­
related cardiac risk in childhood cancer survivors: an analysis of the childhood cancer survivor
study, JCO1801764. J Clin Oncol 2019. https://doi.
org/10.1200/JCO/18/01764
Hoppe BS, Bates JE, Mendenhall NP et al (2019)
The meaningless meaning of mean heart dose
in ­
mediastinal lymphoma in the modern radiation therapy Era. Pract Radiat Oncol pii:s1879–
8500(19)30279–6. https://doi.org/10.1016/j.prro.2019.​
09.015 PMID:31586483
Cutter DJ, Schaapveld M, Darby SC et al (2015) Risk
of valvular heart disease after treatment for Hodgkin
lymphoma. J Natl Cancer Inst 23:107(4)
62. Cella L, Conson M, Caterino M et al (2012) Thyroid
V30 predicts radiation-induced hypothyroidism in
patients treated with sequential chemo-radiotherapy
for Hodgkin’s lymphoma. Int J Radiat Oncol Biol
Phys 82(5):1802–1808
63. Travis LB, Hill DA, Dores GM et al (2003) Breast
cancer following radiotherapy and chemotherapy
among young women with Hodgkin disease. JAMA
290(4):465–475
64. Travis LB, Gospodarowicz M, Curtis RE et al (2002)
Lung cancer following chemotherapy and radiotherapy for Hodgkin’s disease. J Natl Cancer Inst
94(3):182–192
65. Tukenova M, Guibout C, Hawkins M et al (2011)
Radiaiton therapy and late mortality from second
sarcoma, carcinoma, and hematological malignancies
after a solid cancer in childhood. Int J Radiat Oncol
Biol Phys 80(2):339–346
66. Bhatti P, Veiga LH, Ronckers CM et al (2010) Risk of
second primary thyroid cancer after radiotherapy for
a childhood cancer in a large cohort study: an update
from the childhood cancer survivor study. Radiat Res
174(6):741–752
67. Nivolumab and Brentuximab vedotin in treating older
patients with untreated Hodgkin lymphoma. (First
posted May 2, 2016, last updated January 14, 2019).
ClinicalTrials.gov Identifier: NCT02758717
68. Brentuximab vedotin and nivolumab in treating participants with early stage classic Hodgkin lymphoma.
(First posted October 19, 2018, last updated January
16, 2019). ClinicalTrials.gov Identifier: NCT03712202
69. Kumar A, Casulo C, Yahalom J et al (2016)
Brentuximab vedotin and AVD followed by involved-­
site radiotherapy in early stage, unfavorable risk
Hodgkin lymphoma. Blood 128(11):1458–1464
70. Yang JC et al (2018) Brentuximab vedotin and AVD
chemotherapy followed by reduced dose/volume
radiotherapy in patients with early stage, unfavorable
Hodgkin lymphoma. International symposium on
Hodgkin lymphoma (ISHL11), 26–29 Oct, Cologne
71. Duhmke E, Franklin J et al (2001) Low-dose radiation is sufficient for the noninvolved extended-field
treatment in favorable early-stage Hodgkin’s disease:
long-term results of a randomized trial of radiotherapy alone. J Clin Oncol 19:2905–2914
72. Duhmke E, Diehl V et al (1996) Randomized trial
with early-stage Hodgkin’s disease testing 30 Gy vs.
40 Gy extended field radiotherapy alone. Int J Radiat
Oncol Biol Phys 36:305–310
73. Prosnitz LR, Farber LR, Fischer JJ et al (1976)
Long term remissions with combined modality
therapy for advanced Hodgkin’s disease. Cancer
37(6):2826–2833
74. Donaldson SS, Link MP (1987) Combined modality
treatment with low-dose radiation and MOPP chemotherapy for children with Hodgkin’s disease. J Clin
Oncol 5:742–749
75. Dharmarajan KV, Friedman DL, Schwartz CL et al
(2015) Patterns of relapse from a phase 3 study
9
76.
77.
78.
79.
80.
81.
82.
83.
84.
85.
86.
87.
Principles of Radiation Therapy for Hodgkin Lymphoma
of response-based therapy for intermediate-risk
Hodgkin lymphoma (AHOD0031): a report from the
Children’s Oncology Group. Int J Radiat Oncol Biol
Phys 92(1):60–66
Ferme C, Eghbali H et al (2007) Chemotherapy plus
involved-field radiation in early-stage Hodgkin’s disease. N Engl J Med 357:1916–1927
Sasse S, Brockelmann PJ, Goergen H et al (2017)
Long-term follow-up of contemporary treatment in
early-stage Hodgkin lymphoma: updated analyses of
the German Hodgkin study group HD7, HD8, HD10,
and HD11 trials. J Clin Oncol 35(18):1999–2007
Hodgson DC, Koh ES et al (2007) Individualized
estimates of second cancer risks after contemporary
radiation therapy for Hodgkin lymphoma. Cancer
110:2576–2586
Kuttesch JF Jr, Wexler LH et al (1996) Second
malignancies after Ewing’s sarcoma: radiation dose-­
dependency of secondary sarcomas. J Clin Oncol
14:2818–2825
Koontz B, Kirkpatrick J et al (2006) Combined modality therapy versus radiotherapy alone for treatment of
early stage Hodgkin disease: cure versus complications. J Clin Oncol 24:605–611
Salloum E, Doria R et al (1996) Second solid tumors
in patients with Hodgkin’s disease cured after radiation or chemotherapy plus adjuvant low-dose radiation. J Clin Oncol 14:2435–2443
Zhou R, Ng A, Constine LS et al (2016) A comparative evaluation of normal tissue doses for patients
receiving radiaiton therapy for Hodgkin’s lymphoma
on the childhood Cancer survivor study and recent
Children’s oncology group trials. Int J Radiat Oncol
Biol Phys 95(2):707–711
Arakelyan N, Jais J-P, Delwall V et al (2010) Reduced
versus full doses of irradiation after 3 cycles of combined doxorubicin, bleomycin, vinblastine, and dacarbazine in early stage Hodgkin lymphomas. Cancer
116:4054–4062
Tseng YD, Cutter DJ, Plastaras JP et al (2017)
Evidence-based review on the use of proton therapy
in lymphoma from the particle therapy cooperative
group (PTCOG) lymphoma subcommittee. Int J
Radiat Oncol Biol Phys 99(4):825–842
Voong KR, McSpadden K, Pinnix CC et al (2014)
Radiat Oncol 9(94):1–9
Milgrom SA, Chi PCM, Pinnix CC, et al. IJROBP
2017 99(2) S61
Filippi AR, Ragona R, Piva C et al (2015) Optimized
volumetric arc therapy versus 3D-CRT for early
stage mediastinal Hodgkin lymphoma without axil-
88.
89.
90.
91.
92.
93.
94.
95.
96.
97.
197
lary involvement: a comparison of second cancers
and heart disease risk. Int J Radiat Oncol Biol Phys
92(1):161–168
Levis M, De Luca V, Fiandra C et al (2018) Plan optimization for mediastinal radiotherapy: estimation of
coronary arteries motion with ECG-gated cardiac
imaging and creation of compensatory expansion
margins. Radiother Oncol 127(3):481–486
Girinsky T, Pichenot C, Beaudre A et al (2006) Is
intensity-modulated radiotherapy better than conventional radiation treatment and three-dimensional
conformal radiotherapy for mediastinal masses in
patients with Hodgkin’s disease, and is there a role for
beam orientation optimization and dose constraints
assigned to virtual volumes? Int J Radiat Oncol Biol
Phys 64(1):218–226
Petersen PM, Azxnar MC, Berthelsen AK et al (2015)
Acta Oncol 54(1):60–66
Charpentier AM, Conrad T, Sykes J et al (2014)
Active breathing control for patients receiving mediastinal radiaiton therapy for lymphoma: impact
on normal tissue dose. Pract Radiat Oncol 4(3):
174–180
Hoppe BS, Mendenhall NP, Louis D et al (2017)
Comparing breath hold and free breathing during
intensity-modulated radiation therapy and proton
therapy in patients with mediastinal Hodgkin lymphoma. Int J Part Ther 3(4):492–496
Paumier A, Ghalibafian M, Gilmore J et al (2012)
Dosimetric benefits of intensity-modulated radiotherapy combined with the deep-inspiration breath-hold
technique in patients with mediastinal Hodgkin’s lymphoma. Int J Radiat Oncol Biol Phys 82(4):1522–1527
Dabaja BS, Hoppe BS, Plastaras JP et al (2018)
Proton therapy for adults with mediastinal lymphomas: the international lymphoma radiation oncology
group guidelines. Blood 132(16):1635–1646
Hoppe BS, Hill-Kayser CE, Tseng YD et al (2017)
Consolidative proton therapy after chemotherapy
for patients with Hodgkin lymphoma. Ann Oncol
28(9):2179–2184
Nanda R, Flampouri S, Mendenhall NP et al (2017)
Pulmonary toxicity following proton therapy for
thoracic lymphoma. Int J Radiat Oncol Biol Phys
99(2):494–497
Hoppe BS, Flampouri S, Su Z et al (2012) Effective
dose reduction to cardiac structures using protons
compared with 3DCRT and IMRT in mediastinal
Hodgkin lymphoma. Int J Radiat Oncol Biol Phys
84(2):449–455
Principles of Chemotherapy
in Hodgkin Lymphoma
10
David Straus and Mark Hertzberg
Contents
10.1
Historical Introduction 10.2
hemotherapy Applied to Advanced-Stage Hodgkin
C
Lymphoma: Theories and Practice Classes of Active Classical Agents in HL Polychemotherapy: Models and Comparative Clinical Studies MOPP and Derivatives ABVD and Derivatives The Dose/Response Relationship: Norton and Simon Model Sustained/Weekly Regimens Escalated-Dose Regimens High-Dose Treatment and Autologous Stem Cell Transplantation
as Part of Initial Therapy Risk-Adapted Regimens Based on PET Incorporation of Antibody-­Drug Conjugate in Primary Treatment
of Advanced-Stage cHL 10.2.1
10.2.2
10.2.2.1
10.2.2.2
10.2.2.3
10.2.2.4
10.2.2.5
10.2.2.6
10.2.2.7
10.2.2.8
199
200
200
201
201
205
206
207
208
209
210
213
10.3.1
hemotherapy Treatment for Recurrent and Refractory Hodgkin
C
Lymphoma New Systemic Treatments 213
213
10.4
Conclusions 214
10.3
References 214
10.1
D. Straus (*)
Division of Hematologic Malignancies, Department
of Medicine, Memorial Sloan Kettering Cancer
Center, New York, NY, USA
e-mail: strausd@mskcc.org
M. Hertzberg
Department of Haematology, Prince of Wales
Hospital, Randwick, NSW, Australia
e-mail: mark.hertzberg@sesiahs.health.nsw.gov.au
Historical Introduction
Hodgkin lymphoma (HL) was the malignant disease for which the possibility of cure with combination chemotherapy in the majority of patients
was first realized. As such it has provided a model
upon which studies in many other types of malignancy have been based, and it is interesting to follow the trajectory of knowledge from early
single-agent work through combinations,
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_10
199
D. Straus and M. Hertzberg
200
c­ombined modalities, increasing complexity,
and, most recently, selective de-escalation.
Patients with advanced disease represent a minority of those affected by HL. However, these
patients represent the group in which the development and effects of chemotherapy are most
readily appreciated, because the role of radiation
therapy is markedly less than in patients with
localized disease.
As early as 1942, four patients with HL were
treated with nitrogen mustard by Wilkinson and
Fletcher at Manchester Royal Infirmary, although
a military embargo prevented the dissemination
of this information [1]. Similar considerations are
applied to the bombing of the ship “USS Liberty”
on December 3, 1943, in Bari, and the hematological consequences of a nitrogen mustard gas
leak among the survivors. Cornelius Rhoads, an
American cancer researcher, was involved in
their care and understood from his observations
of the effects on the bone marrow and lymphoid
tissue that nitrogen mustard derivatives might be
effective against lymphoid and hematological
malignancies [2, 3]. In 1958, another alkylating
agent, cyclophosphamide, proved effective in
non-Hodgkin lymphoma (NHL) [4]. Shortly after
this, vinblastine was first shown to be an effective
drug in HL, as was vincristine. Although encouraging, the early results of chemotherapy were
modest, with most responses short-lived after
corticosteroids and alkylating and spindle cell
agents [5–7]. There was a prevalent view that
only extensive irradiation could yield complete
cures [8, 9].
One of the first modern randomized studies
was the EORTC H1 trial, which investigated
whether “adjuvant” chemotherapy (weekly
vinblastine for 2 years) could improve the
results over radiotherapy alone [10]. A durable
advantage was seen in the chemotherapy arm
for relapse-free survival (RFS: at 15 years 60%
vs. 38%; P < 0.001) although more than 50%
of patients with mixed-cellularity histology
developed recurrences [11]. To reduce the
relapse rate, irradiation was extended to infradiaphragmatic nodal and spleen areas. Singleagent or doublet chemotherapy was added after
radiotherapy, but no immediate attempt was
made to use polychemotherapy, based upon the
idea that the cure rate would depend upon the
adequacy of irradiation [12, 13]. Two factors
gradually undermined the dominance of strict
pathological delineation and extensive irradiation as the basis of curative therapy in HL: the
advent of accurate cross-sectional imaging by
computed tomographic (CT) scanning and the
recognition that relapses after irradiation alone
had minimal impact on survival owing to the
efficacy of salvage chemotherapy [14]. With
the development of four-drug combination
therapy, which for the first time resulted in
cures for advanced HL without the need for
irradiation, the transition to systemic therapy
began in earnest.
10.2
Chemotherapy Applied
to Advanced-Stage Hodgkin
Lymphoma: Theories
and Practice
10.2.1 C
lasses of Active Classical
Agents in HL (Table 10.1)
Almost every class of chemotherapy drug has
been shown to have some efficacy in HL, with the
possible exception of the antimetabolite drugs
such as 5-fluorouracil [15]. The original combination treatments were based upon evidence of single-agent activity among alkylating agents, vinca
alkaloids, corticosteroids, and the hydralazine
monoamine oxidase inhibitor procarbazine. All of
these produced response rates of over 50% when
used singly in patients not previously exposed to
multiagent chemotherapy (Table 10.1). Later
entrants to this field included the antibiotic drugs
doxorubicin and bleomycin, the nitrosoureas and
dacarbazine, and the podophyllotoxins, all of
which showed appreciable single-agent activity
after prior combination regimens. More recently,
newer cytotoxics such as gemcitabine have been
introduced, often in combination with platinum
drugs, and found to produce significant response
rates in recurrent disease. In 2011, brentuximab
vedotin, an antibody-drug conjugate, was
approved in the USA and conditionally in Europe
10 Principles of Chemotherapy in Hodgkin Lymphoma
Table 10.1 Single-agent activity of cytotoxic drugs in
Hodgkin lymphoma [15]
Complete
Overall
response rate response rate
(%)
(%)
Drug
Single agents tested before combination
chemotherapy
Alkylating agents
Chlorambucil
61
16
Mustine
63
13
Cyclophosphamide
54
12
Vinca alkaloids
Vinblastine
68
30
Vincristine
60
36
Agents mainly tested after prior multiagent
therapy
Dacarbazine
56
6
Nitrosoureas
Carmustine
44
5
Lomustine
48
12
Antibiotics
Doxorubicin
30
5
Bleomycin
38
6
Podophyllotoxin
Etoposide
27
6
Antimetabolite
Gemcitabine
22
0
Antibody-drug conjugate
Brentuximab vedotin 75
34
201
feature of the malignant cells themselves or their
associated inflammatory infiltrate, which may be
critical to sustaining them. The development of
combination therapies has been based mainly
upon the use of agents with non-overlapping toxicity as far as possible, and as cure rates have
risen, the emphasis has fallen increasingly upon
avoiding long-term side effects. The most important among these are infertility and myelodysplasia, mainly caused by the alkylating agents;
pulmonary fibrosis caused by bleomycin and
nitrosoureas; and cardiomyopathy related to
anthracyclines, a risk increased by the concomitant use of mediastinal radiotherapy.
10.2.2 Polychemotherapy: Models
and Comparative Clinical
Studies (Tables 10.2 and 10.3)
10.2.2.1 MOPP and Derivatives
Combination chemotherapy was first attempted
clinically in childhood acute lymphoblastic leukemia by Jean Bernard [18], who designed two
doublets
of
cortisone-methotrexate
and
prednisone-­
vincristine, at the same time as
­pursuing work on chemotherapy for HL. Lacher
and Durant were the first to use doublet combinafor treatment of relapsed or refractory HL after tion chemotherapy in HL with vinblastine and
autologous stem cell transplantation (ASCT) or chlorambucil [19]. At the NCI, Freireich, Frei,
after at least two combination chemotherapy regi- and Katon added 6-mercaptopurine into the more
mens in patients who are not transplant candi- effective VAMP (vincristine, amethopterin, merdates. Approval was granted on the basis of an captopurine, and prednisone) regimen [7]. This
overall response (OR) rate of 75% and a complete led on to MOMP (cyclophosphamide, vincrisresponse (CR) rate of 34% in a phase 2 trial in 102 tine, methotrexate, and prednisone) and MOPP
HL patients relapsed after or refractory to ASCT, (mechlorethamine, vincristine, procarbazine,
response rates approximately twice as high as prednisone), developed by DeVita and Carbone,
those reported for other single agents [16]. This also at the NCI [20, 21]. Some of the critical feaantibody-drug conjugate attaches an anti-CD30 tures of success were prolonged treatment
antibody to a potent antimicrotubular agent, (6 months, more than any other regimen at the
monomethyl auristatin (MMAE), by a protease time); the use of each drug at “optimal” dose and
cleavable linker. MMAE binds to tubulin and dis- schedule with a sliding scale for dose adjustment
rupts the microtubule network, inducing cell cycle according to marrow suppression; an interval of
arrest and apoptosis, a mechanism of action simi- 2 weeks for recovery of normal tissue (marrow,
lar to those for vincristine and vinblastine [17].
GI epithelium), ideally before HL recovery; and
It is clear that HL is broadly sensitive to phase-­ treatment with curative intent rather than palliaspecific, cycle-specific, and non-cycle-specific tion. MOPP provided an 80% response rate and
agents, although it is less clear whether this is a long-term disease-free (DFS) and overall ­survival
D. Straus and M. Hertzberg
202
Table 10.2 Chemotherapy regimens designed for advanced Hodgkin lymphoma
Drugs
Four-drug regimens
MOPP
Mechlorethamine
Vincristine
Procarbazine
Prednisolone
MVPP
Mechlorethamine
Vinblastine
Procarbazine
Prednisolone
ChlVPP
Chlorambucil
Vinblastine
Procarbazine
Prednisolone
COPP
Cyclophosphamide
Vinblastine
Procarbazine
Prednisolone
ABVD
Doxorubicin
Bleomycin
Vinblastine
Dacarbazine
Hybrid regimens
MOPP/ABV
Mechlorethamine
Vincristine
Procarbazine
Prednisolone
Doxorubicin
Bleomycin
Vinblastine
ChlVPP/EVA
Chlorambucil
Vincristine
Procarbazine
Etoposide
Prednisolone
Doxorubicin
Vinblastine
BEACOPP baseline
Bleomycin
Etoposide
Doxorubicin
Cyclophosphamide
Vincristine
Dose, mg/m2
Route
6
1.4 (cap 2 mg)
100
40
Iv
Iv
Po
Po
6
6 (cap 10 mg)
100
40
Iv
Iv
Po
Po
6 (cap 10 mg)
6 (cap 10 mg)
100
40
Po
Iv
Po
Po
650
6
100
40
Iv
Iv
Po
Po
25
10 iu/m2
6
375
Iv
Iv
Iv
Iv
6
1.4
100
40
35
10 iu/m2
6
Iv
Iv
Po
Po
Iv
Iv
Iv
6 (cap 10 mg)
1.4 (cap 2 mg)
90
75
50
50
6 (cap 10 mg)
Po
Iv
Po
Po
Po
Iv
Iv
10 iu/m2
100
25
650
1.4 (cap 2 mg)
Iv
Iv
Iv
Iv
Iv
Schedule
q. 28 days
d1 and 8
d1 and 8
d1–14
d1–14
q. 42 days
d1 and 8
d1 and 8
d1–14
d1–14
q. 28 days
d1–14
d1 and 8
d1–14
d1–14
q. 28 days
d1 and 8
d1 and 8
d1–14
d1–14
q. 28 days
d1 and 15
d1 and 15
d1 and 15
d1 and 15
q. 28 days
d1
d1
d1–7
d1–14
d8
d8
d8
q. 28 days
d1–7
d1
d1–7
d1–5
d1–7
d8
d8
q. 21 days
d8
d1–3
d1
d1
d8
10 Principles of Chemotherapy in Hodgkin Lymphoma
203
Table 10.2 (continued)
Drugs
Procarbazine
Prednisolone
Escalated regimens
Escalated BEACOPP
Bleomycin
Etoposide
Doxorubicin
Cyclophosphamide
Vincristine
Procarbazine
Prednisolone
G-CSF
BEACOPP-14
Bleomycin
Etoposide
Doxorubicin
Cyclophosphamide
Vincristine
Procarbazine
Prednisolone
G-CSF
Weekly regimens
Stanford V
Doxorubicin
Vinblastine
Mechlorethamine
Vincristine
Bleomycin
Etoposide
Prednisolone
VAPEC-B
Doxorubicin
Cyclophosphamide
Etoposide
Vincristine
Bleomycin
Prednisolone
BV/AVD
Brentuximab vedotin
Doxorubicin
Vinblastine
Dacarbazine
Dose, mg/m2
100
40
Route
Po
Po
10 iu/m2
200
35
1250
1.4 (cap 2 mg)
100
40
Iv
Iv
Iv
Iv
Iv
Po
Po
Sc
10 iu/m2
100
25
650
1.4 (cap 2 mg)
100
80
Iv
Iv
Iv
Iv
Iv
Po
Po
Sc
25
6
6
1.4 (cap 2 mg)
5 i.u./m2
60
40
Iv
Iv
Iv
Iv
Iv
Iv
Po
35
350
75–100
1.4 (cap 2 mg)
10
50
Iv
Iv
Iv
Iv
Iv
Po
4-week cycle
d1 and 15
d1 and 15
d1
d8 and 22
d8 and 22
d15 and 16
Daily to week 10 then taper
4-week cycle
d1 and 15
d1
d15–20
d8 and 22
d8 and 22
Daily to week 6 then taper
1.2 mg/kg
25
6
375
Iv
Iv
Iv
Iv
d1 and 15
d1 and 15
d1 and 15
d1 and 15
(OS) of almost 50% and 40%, respectively [22].
The results have held up, and the 20-year analysis
confirmed among 198 patients a CR rate of 81%,
a 19% rate of induction failures, a 36% relapse
rate, and a 54% mortality. Of the 106 deaths, 30
Schedule
d1–7
d1–14
q. 28 days
d8
d1–3
d1
d1
d8
d1–7
d1–14
d8–14
q. 14 days
d8
d1–3
d1
d1
d8
d1–7
d1–7
d8–13
occurred in patients free of disease; among the 92
patients who survived (46%), only two had persistent HL [23]. These results have been reconfirmed in subsequent trials (Table 10.3) [24–27].
Although the rise in cures from HL can be
D. Straus and M. Hertzberg
204
Table 10.3 Summary results of combination chemotherapy regimens used in first-line therapy of advanced
Hodgkin lymphoma
Regimen
MOPP [22–25,
101]
MVPP [38, 102,
103]
ChlVPP [39,
104]
ABVD [24, 47,
68, 69, 77, 78,
105]
MOPP/ABVD
alternating [24,
106, 107]
COPP/ABVD
alternating [62,
108]
MOPP/ABV
hybrid [47, 107,
109, 110]
Stanford V
[66–69]
VAPEC-B [71]
ChlVPP/EVA
[71, 105]
BEACOPP
baseline [62]
Escalated
BEACOPP [62]
BV/AVD [95]
5 year 5 year ≥7 year
CR (%) EFS (%) OS (%) OS (%)
67–81 40–60 65–73 51–70
72–76
60
65–75
57–74
55–60
66
65
68–92
61–80
73–90
77
83–92
65–70
75–84
74
85
69
83
75
80–88
66–75
76–83
72
72–91
54–94
82–96
47
67
62
82–84
79
89
88
76
88
80
81–96
87
91
86
73
82a,b
97a
2 Years
Modified PFS (time to disease progression, death, or
modified progression)
a
b
ascribed to multiple advances and not just the
introduction of effective chemotherapy, the 1970
report convinced almost all groups treating HL to
accept the inclusion of polychemotherapy
(MOPP or MOPP derivatives) in the treatment
strategy for localized as well as advanced disease. In almost all instances where a combined
treatment was compared to irradiation alone,
whether patients were staged or not with laparotomy, advantages in terms of response and disease- and relapse-free survival were observed
when MOPP or a MOPP-derived chemotherapy
was used [28].
Analysis of the results with MOPP has proven
a fruitful source of information to design and
interpret future studies. Thus, complete response
was seen to be a prerequisite for sustained remission, and a high percentage of complete responses
was correlated with higher survival rates. Capping
the vincristine dose at 2 mg may have been detrimental to the results. Patient and initial disease
characteristics were good predictors of outcome,
with confirmation of the adverse prognostic significance of systemic “B” symptoms.
Maintenance treatment with intermittent MOPP
or carmustine did not appear beneficial [29]. In
patients treated previously by irradiation and
chemotherapy, MOPP was less well-tolerated
and less effective [30]. Conversely, retreatment in
relapsed patients whose initial remission lasted
over a year proved efficient on the second occasion [31]. MOPP therapy carries consequences in
terms of carcinogenicity, in particular with secondary acute myeloid leukemia (AML) [32, 33].
It is also responsible for impaired fertility in both
men and women [34]. Immunosuppression
related to the treatment, or to the underlying disease, brings risks of different types, in particular
of opportunistic infection [35].
There were many attempts to improve upon
these results. The three best-known MOPP-­
derived regimens have been MVPP, with vinblastine instead of vincristine; ChlVPP (chlorambucil,
vinblastine, procarbazine, and prednisolone); and
COPP, with an additional substitution of mechlorethamine, replaced by chlorambucil or cyclophosphamide (Tables 10.2 and 10.3). These
alternatives have never undergone direct comparison, and historical controls are difficult to
interpret. In addition, the proportion of patients
who have also had radiotherapy varies considerably between series. For example, in the NCI
series, 32/198 patients had been irradiated prior
to MOPP, and 28/198 patients received total
nodal irradiation (TNI) “to prevent recurrent disease in previously involved nodes” as consolidation after chemotherapy. MVPP, devised in the
UK, proved easier to handle than MOPP (with
less constipation and neurological toxicity), but
was slightly more myelotoxic [36–38]. ChlVPP
appeared more patient-friendly, inducing minimal nausea/vomiting, constipation or neurologic
toxicity, and limited hematotoxicity, and the
number of cycles could be adapted to the
10 Principles of Chemotherapy in Hodgkin Lymphoma
response: a maximum of five beyond CR. The
66% OS rate in advanced HL was comparable to
mustine-containing regimens, at lower toxic cost,
for all of these acute toxicities, except myelosuppression [39, 40]. COPP is less myelotoxic than
MOPP and is often used in children [41].
205
HL. In 232 patients, a combined modality
approach of three cycles before and after extensive irradiation yielded a CR rate of 80.7% after
MOPP/radiotherapy and 92.4% after ABVD/
radiotherapy (P < 0.02). At 7 years follow-up,
ABVD surpassed MOPP for freedom from progression (FFP) (80.8% vs. 62.8%; P < 0.002),
10.2.2.2 ABVD and Derivatives
RFS (87.7% vs. 77.2%; P = 0.06), and OS (77.4%
The ABVD regimen (doxorubicin, bleomycin, vs. 67.9%; P = 0.03). With longer follow-up, the
vinblastine, and dacarbazine) was devised just disadvantages of MOPP in terms of fertility dam10 years after MOPP, in 1973, for intravenous-­ age and second myelodysplasia (MDS) and leuonly administration at fixed 2-week intervals. kemia were also more apparent. The final
Like MOPP, ABVD was a combination of hema- establishment of ABVD as the favored regimen,
totoxic and neurotoxic drugs. Both doxorubicin at least in North America, was based on two ranand vinblastine had been shown highly effective domized trials for advanced HL. In the first,
in HL. The results with dacarbazine were numer- MOPP vs. ABVD vs. MOPP alternating with
ous but possibly less convincing, and bleomycin ABVD were associated with 5-year failure-free
was also felt to have considerable potential [10, survival rates of 50%, 61%, and 65%, respec42–45]. By comparison to MOPP, hematotoxicity tively. There was less toxicity with ABVD than
after ABVD was predictable, noncumulative, and with MOPP or MOPP alternating with ABVD
milder as a result of the intravenous dosing and and no significant difference in survival among
short intervals. Further, ABVD was far less neu- the three regimens [24]. A second trial compared
rotoxic. Bonadonna developed ABVD at the ABVD with a hybrid regimen, MOPP/ABV;
Milan NCI with the intention: “to compare the 5-year failure-free survival rates were 63% and
efficacy of ABVD with MOPP, and to demon- 66%, respectively. There was a greater incidence
strate absence of cross-resistance between the of acute toxicity, myelodysplastic syndrome, and
two regimens” [46]. The results of MOPP were leukemia for MOPP/ABV compared with
well established and the potential of ABVD in ABVD. Again, there was no significant differterms of “alternative to MOPP to be used either ence in survival [47].
in MOPP failures or in sequential combination
Currently, ABVD is considered by most inveswith MOPP” was clearly in the mind of the tigators as the standard chemotherapy for most
authors, based on these very early results achieved patients with HL, with the possible exception of
in 45 patients. No significant cardiac toxicity was high-risk patients with advanced disease and
seen in this first series, probably because of the poor prognostic features. Reasons to avoid
relatively small cumulative dose of doxorubicin ABVD relate to previous lung impairment and
(6 cycles = 300 mg/m2), the short follow-up, and decreased left ventricular ejection fraction.
the small numbers of patients. Conversely, bleo- Hematological toxicity is usually moderate, and
mycin pulmonary toxicity was apparent from the ABVD may be delivered safely at full dose and
outset, while the effects upon fertility were ini- on schedule to a non-selected average population
tially overestimated through short observation of adult patients without the need to modify doses
which did not take into account the reversal of in the presence of neutropenia [48]. The most fretemporary amenorrhea in some women.
quent serious toxicity with ABVD is pulmonary
It took a surprisingly long time for ABVD to fibrosis, which may be fatal [49]. The discontinube accepted as a standard of care, and it was ini- ation of bleomycin for toxicity during ABVD
tially considered only as a salvage treatment in treatment does not appear to have an adverse
MOPP failures. However, the Milan group under- effect on outcome, which calls into question the
took a larger trial, comparing MOPP and ABVD importance of bleomycin in the ABVD regimen
directly in patients with stage IIB, IIIA, and IIIB [47, 49–51]. This possibility has recently been
206
tested prospectively in a randomized study of
patients showing a good early response to ABVD,
where patients either continued all drugs, or AVD
only [17, 52]. The results confirmed the excess
toxicity associated with bleomycin, particularly
reduced lung function and more instances of
venous thromboembolism. There was no decrease
in efficacy by omission of bleomycin from the
last four cycles of treatment.
10.2.2.3
The Dose/Response
Relationship: Norton
and Simon Model
Much of the thinking about how to maximize
the cure rate in lymphoma has centered upon the
relationship between dose and response to cytotoxic therapy. Theories of tumor cell ecology
have suggested that as the mass of disease is
reduced, the growth fraction may rise. This,
together with the assumed selection of resistant
subclones, underlies the idea that tumor eradication is dependent upon the delivery of treatment
at adequate dose intensity early in a course of
treatment. If doses are too small or too
infrequent, the fractional cell kill might be
­
expected to decline and allow the emergence of
resistance [53].
Three prospective clinical trials have directly
addressed the question of dose versus response
using the same chemotherapy drugs in both
arms. In the first-line treatment of advanced disease, a critical study, HD9, was performed by the
German Hodgkin Study Group (GHSG), as
detailed later, in which patients were randomized between the baseline BEACOPP (bleomycin, etoposide, doxorubicin, cyclophosphamide,
vincristine, procarbazine, and prednisone) regimen and an escalated regimen, with the doses of
doxorubicin, cyclophosphamide, and etoposide
increased to 140%, 185%, and 200%, respectively. This resulted in an increase in freedom
from treatment failure (FFTF) at 5 years from
76% to 87% (P < 0.01), which was translated
into a small but significant improvement in survival on longer follow-up (80% vs. 86% at
10 years; P = 0.0053). This was at the cost of an
increased risk of MDS and AML in the escalated
arm, but at a frequency too low to reverse the
D. Straus and M. Hertzberg
gain in survival from better control of the
­lymphoma [54].
There are two randomized studies for recurrent disease which have yielded similar data on
the dose-response relationship. The UK group
compared the myeloablative BEAM (carmustine,
etoposide, cytarabine, and melphalan) regimen to
mini-BEAM, which uses the same drugs at nonmyeloablative doses. The high-dose treatment
yielded superior PFS (P = 0.005), although the
trial was closed with only 44 patients recruited
and had insufficient power to demonstrate a survival advantage [55]. A study of similar design
was conducted by the GHSG, and this too demonstrated superior FFTF at 3 years (55% for
BEAM, 34% for nonmyeloablative dexa-BEAM,
P = 0.019), although once again no survival difference could be demonstrated [56].
While there is good evidence for an overall
dose-response relationship, there are several
areas of continuing uncertainty. For example, it is
not clear whether the dose of treatment over a
whole course is the critical determinant of outcome, or whether initial dose intensity during the
first weeks of treatment is more important. From
retrospective analyses comparing outcomes to
doses administered, it appears that the most influential factor is the total dose of treatment given,
with some scope for compensating suboptimal
early treatment by later escalation, a finding that
may distinguish HL from many other malignancies [57–59].
Dose/Response Relationships
and Treatment Tolerance: An Individual
Characteristic?
A dose response for both malignant and normal
tissue toxicity is well-recognized, raising the
question of whether the efficacy of tumor control can be related to toxic side effects, effectively using each subject as his or her own
pharmacodynamic control. The GHSG explored
hematotoxicity as a surrogate for pharmacological and metabolic heterogeneity, in relation to
reduced systemic dose and disease control.
Patients treated with various regimens in the
HD6 trial (validated on two other cohorts) were
retrospectively classified as showing WHO
10 Principles of Chemotherapy in Hodgkin Lymphoma
grade l­
eukocytopenia of 0–2 and >2,
­respectively. Patients with a high hematological
toxicity had a 5-year FFTF rate of 68% versus
47% for those with low toxicity, independent of
the actual drug doses received [60]. No pretreatment pharmacokinetic parameters could be
found to explain these observations; however,
recent work from the French Study Group of the
Adult Lymphoma (GELA) has explored polymorphisms in a population of HL patients that
might determine anticancer agent metabolism.
The UGT1A1 polymorphism has been identified as a possible candidate for influencing the
metabolism of several anticancer drugs and
patient outcomes [61]. Unfortunately, similar
dose-response relationships are also seen for
long-term toxicities, for example, infertility and
secondary leukemias [62–64].
10.2.2.4 Sustained/Weekly Regimens
Pursuing the idea of increased dose intensity, several groups developed novel, brief duration regimens for the treatment of advanced HL. The
rationale for the development of these regimens
was, firstly, increased dose intensity of chemotherapy by reduction in the total duration of treatment but an increase in the number of different
agents and, secondly, reduced cumulative doses
of drugs responsible for long-term toxic effects,
including alkylating agents, doxorubicin, and
bleomycin. The PACEBOM, VAPEC-B, and
Stanford V regimens were all designed to deliver
weekly treatments, alternating between myelosuppressive and nonmyelosuppressive agents.
The preliminary results from single-arm studies
appeared promising, with high response and survival rates [65]. Unfortunately, the results of randomized trials did not confirm the early promise
of these regimens.
The Stanford V program developed from the
close collaboration of radiotherapy and chemotherapy, endeavoring to minimize the use of each
modality to achieve improved results with less
toxicity. Initial chemotherapy was composed of
the standard drugs from the MOPP/ABVD
scheme (mechlorethamine, doxorubicin, bleomycin), plus etoposide, with dose intensity increased
for better and earlier tumor response, while
207
cumulative doses, thought to be responsible for
late toxicity (marrow, heart, lung), were reduced.
The use of alkylating agents was limited in order
to avert gonadal damage. The final scheme was
an abbreviated 12-week program with radiotherapy started 2–4 weeks after chemotherapy,
restricted to sites at higher risk for relapse (bulky
sites), and delivered at 36 Gy, in order to reduce
the incidence of late cardiopulmonary effects,
and “mini-mantle” instead of mantle fields, sparing the axillae to decrease the risk of secondary
breast carcinoma. The results of the initial
Stanford V phase 2 approach were confirmed in
the Eastern Cooperative Oncology Group
(ECOG) E1492 study in 45 patients, of whom
87% received radiotherapy; FFP was 85% at
5 years, and OS was 96% with one death from
HL and one from an M5 AML [66]. Later analysis confirmed these excellent results and the relative preservation of fertility in both women and
men; no case of secondary MDS/leukemia or
NHL had been registered at a 65 months median
follow-up [67].
A randomized trial (Italian Lymphoma Group:
ILL) compared Stanford V to mechlorethamine,
vincristine, procarbazine, prednisone, epidoxorubicin, bleomycin, vinblastine, lomustine, doxorubicin, and vindesine (MOPPEBVCAD) and to
ABVD as the standard in 355 patients with stage
IIB–IV HL. In this trial, the Stanford V arm was
inferior to the other two arms in terms of 5-year
FFS (54% vs. 78% for ABVD and 81% for
MOPPEBVCAD, respectively (P < 0.01) for
comparison of Stanford V with the other two regimens) [68]. However, only 66% of patients in
the Stanford V arm received irradiation, against
87% in the ECOG phase 2 study: this is important in a strategy that was originally designed to
combine both modalities. The Stanford V program was also compared to ABVD in a large prospective trial run by the UK National Cancer
Research Institute Lymphoma Group (NCRI) in
520 patients with stage IIB–IV HL. Results in the
Stanford V and in the ABVD arm were similar
for 5-year PFS and OS rates (76% and 90%, for
ABVD; 74% and 92% for Stanford V, with radiotherapy administered in 53% and 73%, respectively) [69]. The North American Intergroup trial
208
led by ECOG (E2496) compared ABVD with
IFRT only to bulky mediastinal sites with the
combined modality Stanford V. There was no difference in response rates or in 5-year FFS or OS
rates between the two arms of the trial. The relatively extensive use of radiotherapy required to
achieve optimum results for weekly regimens
makes them a less attractive choice for many
patients: in the UK study, 73% of patients treated
with Stanford V received consolidation radiotherapy, compared to 37% in the previous UK
study using ABVD in a similar group of patients.
In E2496, 75% of patients on the Stanford V regimen received radiation therapy, while 41% of
those on ABVD had irradiation of bulky mediastinal sites [70]. The short 12-week duration of the
Stanford V regimen has some appeal for patients
and remains a reasonable approach for those with
low-risk non-bulky disease, for whom limited or
no irradiation is needed, but this is only a
minority.
The only other weekly regimen to be compared with a hybrid regimen in a randomized trial
featured myelosuppressive (doxorubicin, cyclophosphamide, and etoposide) and relatively nonmyelosuppressive (vincristine and bleomycin)
drugs given on an alternating weekly basis for
11 weeks: VAPEC-B. This regimen was compared to a hybrid ChlVPP-EVA schedule for
advanced disease, was expected to still be significantly more myelosuppressive and to impair fertility, and showed inferior PFS for the weekly
regimen in all but the best prognosis subgroup.
Event-free survival at 5 years in newly diagnosed
patients with advanced disease following the
hybrid regimen was 78% versus 58% for
VAPEC-B, which translated into better OS, at
89% versus 79% [71].
10.2.2.5 Escalated-Dose Regimens
In order to spare patients the acute gastrointestinal
and hematologic toxicities, the original recommendation of the NCI to follow a “sliding scale”
of dose adaptation for MOPP was gradually
superseded by fixed doses at well-tolerated levels
and intervals. Retrospective studies of MOPP and
MVPP suggested that the cumulative dose, as
much as frequency of administration or dose
D. Straus and M. Hertzberg
intensity, might determine the outcomes [25, 72].
These observations also appear to hold for ABVD
[59], although all these studies are retrospective
and need to be confirmed in a prospective study.
The GHSG has pioneered the exploration of
two levels of dose increment, in the conventional
dose range, by reducing the length of treatment
and adding etoposide to the standard regimen,
COPP/ABVD [73]. Further intensification was
carried out by increasing the myelosuppressive
drug doses, with growth factor support. Both
intensified regimens provided higher CR and
FFTF and, crucially, statistically higher OS rates
as compared to standard COPP/ABVD [54]. The
early effects of dose intensification were maintained in the long-term results at 10 years: FFTF
was 64%, 70%, and 82% with OS rates of 75%,
80%, and 86% for patients treated with standard
COPP/ABVD, BEACOPP baseline, and
BEACOPP escalated, respectively (P < 0.001)
[62]. The higher overall chemotherapy doses, as
given in the escalated BEACOPP scheme, appear
to provide greater disease control than any of the
previous or contemporary regimens. This is supported by the very low number of deaths due to
the progression of lymphoma (2.8%). The GHSG
has conducted a series of studies, HD12, HD15,
and HD18, all using escalated BEACOPP in
advanced HL patients (under the age of 61)
whose results replicate closely those of the escalated BEACOPP arm in the HD9 study [74–76].
The GHSG reported early on its concerns for
the immediate toxicity, especially among
patients older than 65, and, in younger patients,
impaired fertility and risk of MDS or secondary
AML. A review of the HD9 results concerning
the cumulative incidence of all second tumors at
10 years confirmed that the rate for AML/MDS
was lower after COPP/ABVD (0.4%) versus
BEACOPP baseline (2.2%) and BEACOPP
escalated (3.2%; log-rank test; P = 0.03).
However, counting all secondary malignancies,
there was no difference (5.3% after COPP/
ABVD, 7.9% after BEACOPP baseline, and
6.5% after BEACOPP escalated) [62].
The immediate and long-term toxic effects of
escalated BEACOPP and the reluctance of many
specialists to consider COPP/ABVD as a ­standard
10 Principles of Chemotherapy in Hodgkin Lymphoma
209
comparator have hindered acceptance of escalated BEACOPP as a new standard of care. Two
Italian trials, HD2000 and GSM-HD, have demonstrated superior PFS with escalated BEACOPP
in comparison to ABVD. In HD2000, BEACOPP
resulted in an 81% (95% CI, 70–89%) 5-year
PFS versus 68% (95% CI, 56–78%) for ABVD,
but no significant OS difference was observed
[77]. Similarly, the GSM-HD trial demonstrated
a higher 3-year FFP for escalated plus baseline
BEACOPP (4 + 4) versus ABVD (87 ± 3% and
71 ± 4%), respectively, but freedom from second
progression (FF2P) and OS were alike [78].
ABVD was declared preferable, taking into
account the lesser toxicity, including fewer toxic
deaths (one vs. six).
The outstanding results of escalated
BEACOPP, despite the toxicity, have made it
most appealing for high-risk patients. This has
been called into question by results in two recent
randomized clinical trials. In a multi-institutional
Italian trial comparing ABVD with BEACOPP
(4 cycles escalated dose + 4 cycles standard dose)
for patients with stages IIB, III, or IV HL, the
superior freedom from first progression for
BEACOPP was confirmed (at 7 years, 73% for
ABVD vs. 85% for BEACOPP; P = 0.004),
which was the primary endpoint of the trial.
However, there was no significant difference in
freedom from second relapse following ASCT or
in OS between the two treatment arms. The
treatment-­
related mortality was 4% for
BEACOPP vs. 1% for ABVD [79]. This suggests
that most patients can be treated initially with
ABVD and only those who relapse be salvaged
with ASCT and thus exposed to a treatment-­
related mortality similar to that with initial
BEACOPP treatment. The EORTC randomized
patients with high-risk stages III or IV HL (international prognostic score ≥ 3) to BEACOPP
(4 cycles dose escalated + 4 cycles standard dose)
or ABVD. There was no significant difference in
4-year event-free survival (EFS) or OS, which
was the primary endpoint, although this trial also
confirmed a superior PFS for BEACOPP [80].
Progression-free survival may not be the most
clinically important treatment result, and these
two trials suggest that ABVD is an acceptable
initial treatment approach even for high-risk
advanced-stage HL patients because of the effectiveness of salvage ASCT in the minority of
patients who relapse.
As with ABVD, it was found that omission of
bleomycin because of toxicity during treatment
with BEACOPP did not have an adverse impact
on PFS or OS. In addition, with this intensive
regimen, omission of vincristine during treatment because of toxicity also had no adverse
impact on these outcomes [81].
10.2.2.6
High-Dose Treatment
and Autologous Stem Cell
Transplantation as Part
of Initial Therapy
Attempts have been made to improve results by
using intensified consolidation and peripheral
blood stem cell (PBSC) rescue for patients considered at high risk. Three randomized studies
have explored this concept for HL. The Scotland
and Newcastle Lymphoma Group HD3 study
randomized 65 out of 126 high-risk patients,
resulting in a nonsignificant advantage for the
conventional arm (time to treatment failure 85%
vs. 79%; P = 0.35) [82]. A European study of
similar design randomized 163 high-risk patients
achieving CR or partial response (PR) after four
cycles of ABVD or an equivalent regimen to
receive high-dose therapy plus ASCT (83
patients) or four more cycles of conventional chemotherapy (80 patients). There was no evidence
of a benefit to the group receiving high-dose therapy (CR 92% vs. 89%, 5-year FFS 75% vs. 82%,
and OS 88% vs. 88%, respectively) [83].
The Groupe Ouest-Est d’Etude des Leucémies
et Autres Maladies du Sang (GOELAMS) undertook a randomized study in 158 high-risk patients,
comparing conventional intensive chemotherapy
with vindesine (5 mg/m2), doxorubicin (99 mg/
m2), carmustine (140 mg/m2), etoposide (600 mg/
m2), and methylprednisolone (600 mg/m2)
(VABEM) followed by low-dose lymph node
irradiation in 82 patients versus four cycles of
ABVD followed by myeloablative carmustine
(300 mg/m2), etoposide (800 mg/m2), cytarabine
(1600 mg/m2), and melphalan (140 mg/m2) and
ASCT in 76 patients. The results were ­remarkably
D. Straus and M. Hertzberg
210
similar for CR (89% vs. 88%), 5-year FFTF (79%
vs. 75%), and OS (87% vs. 86%) [84].
In summary, there is no evidence to support
the use of high-dose consolidation at first remission in HL at present.
when compared with addition of IFRT; OS did
not differ between groups, as the 22 patients who
relapsed without further IFRT were successfully
treated with salvage therapy. Of note, 5 of the 22
received only radiation therapy as a salvage treatment, and only 7 of 22 received chemotherapy
10.2.2.7 Risk-Adapted Regimens
followed by ASCT. Negative PET was defined as
Based on PET
a Deauville score of 1–2 (FDG uptake less than
Response to treatment for classical HL (cHL) is mediastinal blood pool).
assessed by positron emission tomography–
Another phase 2 trial confirmed an excellent
computed tomography (PET/CT) at end of treat- PFS for most patients treated with a short
ment (EOT). Negative PET/CT is associated course of ABVD alone. CALGB 50604 treated
with a 10% or lower likelihood of relapse [85]. patients with stages I/II non-bulky cHL with
Interim PET/CT after one or two cycles of two cycles of ABVD. Interim PET/CT was perABVD or similar regimens is also highly predic- formed and centrally reviewed. Patients whose
tive of outcome [86, 87].
interim PET/CT was negative, defined as
Three recent clinical trials have utilized PET/ Deauville scores of 1–3 (FDG uptake less than
CT to determine if negative PET after or during liver), received two more cycles of ABVD (total
treatment will identify a population of early-­ four cycles) and no irradiation (135/149; 91%).
stage patients with non-bulky disease who can Patients whose interim PET/CT was positive
safely be treated with ABVD alone (Table 10.4). received two cycles of more intensive chemoThe Randomized Phase 3 Trial to Determine the therapy with escalated BEACOPP and IFRT to
Role of FDG-PET Imaging in Clinical Stages IA/ a dose of 3060 cGy (13/149; 9%). Estimated
IIA Hodgkin lymphoma (RAPID) found high PFS was 91% at 3 years for the interim PETrates of 3-year PFS among patients who were negative group. The estimated 3-year PFS for
PET-negative after three cycles of ABVD, regard- the interim PET-positive group was signifiless of whether they received IFRT or no further cantly lower, at 66%, than for the interim PETtreatment (90.8% vs. 94.6%; P = 0.16) [88]. negative group (P = 0.011), s­ uggesting that the
Though the PFS rate for chemotherapy alone was intensive treatment regimen did not provide
excellent, non-inferiority criteria were not met benefit [89].
Table 10.4 Salvage regimens in common use for recurrent/refractory Hodgkin lymphoma drugs
Dexa-BEAM
Dexamethasone
Carmustine
Etoposide
Cytarabine
Melphalan
DHAP
Dexamethasone
Cytarabine
Cisplatin
ESHAP
Etoposide
Cytarabine
Cisplatin
Methylprednisolone
Dose, mg/m2
Route
24 mg daily
60
250
100 bd
20
Po
Iv
Iv
Iv
Iv
40 mg daily
2000 bd
100
Iv
Iv
Ivi
40
2000
25
500 mg daily
Iv
Iv
Ivi
Iv
Schedule
q. 21d
d1–10
d2
d4–7
d4–7
d3
q. 21d
d1–4
d2
d1
q. 21d
d1–4
d5
d1–4
d1–5
10 Principles of Chemotherapy in Hodgkin Lymphoma
211
Table 10.4 (continued)
ICE
Ifosfamide
Carboplatin
Etoposide
GDP
Gemcitabine
Dexamethasone
Cisplatin
GVD
Gemcitabine
Vinorelbine
Liposomal doxorubicin
IGEV
Ifosfamide
Gemcitabine
Vinorelbine
Prednisone
BeGEV
Bendamustine
Gemcitabine
Vinorelbine
Prednisone
BV-Benda
Bendamustine
Brentuximab vedotin
BV-ESHAP
Brentuximab vedotin
Etoposide
Cytarabine
Cisplatin
Methylprednisolone
BV-DHAP
Brentuximab vedotin
Dexamethasone
Cytarabine
Cisplatin
Dose, mg/m2
Route
5000
AUC 5
100
Ivi
Iv
Iv
1000
40 mg daily
75
Iv
Po
Iv
Schedule
q. 21d
d2
d2
d1–3
q. 21d
d1 and 8
d1–4
d1
1000
20
15
Iv
Iv
Iv
d1 and 8
d1 and 8
d1 and 8
2000
800
20
100
Iv
Iv
Iv
Po
d1–4
d1 and 4
d1 and 4
d1–4
90
800
20
100
Iv
Iv
Iv
Po
d2 and 3
d1 and 4
d1
d1–4
90
1.8 mg/kg
Iv
Iv
d1 and 2
d1
0.9–1.2–1.8 mg/kg
40
2000
25
500 mg daily
Iv
Iv
Iv
Iv
Iv
d1–4
d5
d1–4
d1–5
1.8 mg/kg
40 mg daily
2000 bd
100
Iv
Iv
Iv
Iv
d1
d1–4
d2
d1
The larger phase 3 H10 trial compared a similar interim PET-adapted approach to combined-­
modality therapy (CMT) for all patients [90].
Patients with early-stage cHL received two cycles
of ABVD and underwent interim PET/
CT. Interim PET-negative favorable patients in
the PET-adapted arm received two more cycles of
ABVD (total four) and no RT, while those in the
CMT arm received one more cycle of ABVD
(total three) and involved-node radiation therapy
(INRT). Favorable patients who were interim
PET-positive in the PET-adapted arm received
two cycles of escalated BEACOPP and INRT,
while those in the CMT arm received two more
cycles of ABVD and INRT. Interim PET-negative
unfavorable patients in the PET-adapted arm
received four more cycles of ABVD (total six)
and no RT, while those who were interim PETnegative in the CMT arm received two more
cycles of ABVD (total four) and INRT. Interim
PET-positive unfavorable patients in the PET-­
adapted arm received two cycles of escalated
BEACOPP and INRT, while interim PET-positive
patients in the CMT arm received two more
212
cycles of ABVD (total four) and INRT. Overall,
in favorable and unfavorable groups together, this
trial demonstrated a 5-year PFS benefit for a
CMT regimen as compared with a PET-adapted
regimen (91% vs. 77%; P = 0.002) [91]. The
5-year PFS for interim PET-negative patients in
the favorable group was 87% for the PET-adapted
arm versus 99% for the CMT arm. The 5-year
PFS for interim PET-negative unfavorable
patients was 90% for the PET-adapted arm versus
92% for the CMT arm. PFS favored CMT over
PET-adapted treatment, and non-inferiority could
not be ­demonstrated in the large number patients
undergoing analysis.
In the recent HD16 randomized trial, early-­
stage favorable cHL patients received two cycles
of ABVD + 20 Gy IFRT (control) or two cycles
of ABVD plus PET, with PET-negative patients
receiving no further treatment and PET-positive
patients receiving 20 Gy IFRT (risk adapted).
Following two cycles of ABVD only, PET-­
negative patients had a relapse rate of 10% at
5 year PFS, higher than those who received
standard two cycle ABVD + 20 Gy IFRT [92].
These trials demonstrate a 5–10% higher relapse
rate for 2–4 cycles of ABVD alone as compared
with CMT for favorable early-stage cHL. Opinions
differ as to whether it is more important to reduce
the late risks of radiotherapy with chemotherapy
only, given the excellent salvage options, or to provide a more optimal PFS with frontline treatment
by adding radiotherapy to chemotherapy.
Four recent trials have employed interim PET
after two cycles of chemotherapy to tailor treatment for patients with advanced-stage
cHL. S0816, a phase 2 trial conducted by the US
Intergroup, treated stage III and IV patients with
two cycles of ABVD followed by interim PET/
CT. Interim PET-negative patients received 4
more cycles of ABVD, while those who were
interim PET-positive received two cycles of escalated BEACOPP. The estimated 2-year PFS was
82% for interim PET-negative patients and 64%
for interim PET-positive patients. Of note, there
were two treatment-related deaths (4%) among
D. Straus and M. Hertzberg
the 49 interim PET-positive patients who receive
escalated BEACOPP [93].
The Response-Adapted Trial in Advanced
Hodgkin Lymphoma (RATHL) treated patients
with stages IIB, III, and IV and high-risk stage IIA
with two cycles of ABVD followed by interim
PET/CT. Patients who were interim PET-negative
were randomized to treatment with four cycles of
ABVD or four cycles of AVD without bleomycin.
Patients who were interim PET-positive were
treated with escalated BEACOPP or BEACOPP-14
depending on results of further interim PET/CT
studies. For post-cycle 2 interim PET-negative
patients, the 3-year PFS was 85.7% for the ABVD
and 84.4% for the ABVD/AVD groups, respectively. For the interim PET-positive patients treated
with BEACOPP, the 3-year PFS was 67.5%. These
findings justify reducing exposure to bleomycin
with its attendant pulmonary toxicity for patients
with advanced-stage cHL who are interim PETnegative after two cycles of ABVD [52].
The HD18 trial administered two cycles of escalated BEACOPP to patients with advanced-­stage
cHL followed by interim PET. Patients who were
interim PET-negative just received two more cycles
of escalated BEACOPP and no additional radiotherapy. PET-positive patients after two cycles of
escalated BEACOPP received a total of four or six
additional cycles and radiotherapy to PET-positive
residual disease. For PET-negative patients, 5-year
PFS was 91.2% for 8/6 escalated BEACOPP and
91.8% for four escalated BEACOPP, and there was
less toxicity in the latter group [94].
The LYSA AHL2011 trial randomized
advanced-stage cHL patients to standard treatment with six cycles of escalated BEACOPP plus
interim PET after two and four cycles or experimental treatment. In the experimental arm, treatment was initiated with two cycles of escalated
BEACOPP. Following interim PET/CT, treatment
was changed to four cycles of ABVD in interim
PET-negative patients, while interim PET-positive
patients continued four cycles of escalated
BEACOPP. There was no significant difference in
4-year PFS between the standard (86.2%) and
10 Principles of Chemotherapy in Hodgkin Lymphoma
213
experimental arms (85.7%). These results suggest
that treatment can be safely de-­escalated to ABVD
for patients with advanced-­stage disease who are
PET-negative after two cycles of escalated
BEACOPP (Casanovas O, Presentation USHL11,
Cologne, October 29, 2018).
The goal of a PET-adapted approach, by starting
with either BEACOPP escalated or ABVD, is to
maintain efficacy and minimize long-term toxicities. Ideally, a more effective risk-allocation strategy would use novel biomarkers such as TARC or
ctDNA, with or without baseline PET parameters
such as total metabolic tumor volume (TMTV).
Such a strategy would allow patients with highestrisk baseline features to receive the potential benefit of more-intensive initial therapy and would
identify those for whom less-intensive or de-­
escalation strategies can be successfully applied.
ment for all patients with stages III and IV cHL
or in particular subgroups [95].
10.2.2.8
Incorporation of Antibody-­
Drug Conjugate in Primary
Treatment of AdvancedStage cHL
Brentuximab vedotin (BV) is an anti-CD30drug conjugate consisting a monoclonal antibody to CD30 linked to monomethyl auristatin
E, a tubulin inhibitor. As a single agent in
relapsed/refractory cHL, it achieves an overall
response rate of 75% and a CR rate of 34%,
which is at least twice as effective as any single
conventional chemotherapy agent [16]. The
ECHELON-1 trial was an open-label, multicenter randomized phase 3 trial of six cycles of
standard ABVD versus six cycles of BV + AVD
in newly diagnosed patients with stages III and
IV cHL. The 2-year modified PFS was 77.2%
for ABVD and 82.1% for BV + AVD (P = 0.03).
This result led to approval of BV + AVD for this
indication by the US Food and Drug
Administration. Some subgroups seemed to particularly benefit from BV + AVD, and further
analyses are being performed to better define
these groups. Given cost and toxicity considerations, it is unclear, at least in the USA, whether
BV + AVD will be adopted as a standard treat-
10.3
Chemotherapy Treatment
for Recurrent and Refractory
Hodgkin Lymphoma
10.3.1 New Systemic Treatments
There have been relatively few new conventional
cytotoxic agents developed recently for HL, but
both monoclonal antibodies, immune therapies,
and small molecule therapeutics targeting specific abnormal pathways in HL have shown some
promising results.
Antibody therapies have been directed at relatively specific molecules, such as CD30 on the
surface of Reed-Sternberg cells, but the results
with an unconjugated anti-CD30 were discouraging, probably because it targets only a small proportion of the cells within a mass of lymphoma
[96]. On the other hand, antibody-drug conjugates
(ADC) have shown very promising results, with a
response rate of 75% reported using brentuximab
vedotin for patients with recurrent and refractory
disease, as described in Sect. 10.2.1 [16].
Anti-CD20, given with the intention of targeting the infiltrating B cells and interrupting autocrine growth factor loops, has shown some
promise in an early pilot study [97], but awaits
confirmatory data from a prospective trial. This
approach may find more application in the treatment of nodular lymphocyte predominant disease, in which CD20 is present on the surface of
the malignant cells [98].
Among the small molecule therapies being
tested, proteosome inhibitors have been disappointing in HL [99], whereas inhibitors of histone deacetylase (HDACi) have resulted in
significant responses in early-phase studies,
despite significant marrow toxicity [100]. It is not
clear whether the principal target of HDACis is
the malignant cell itself or the surrounding
inflammatory infiltrate, but further studies using
D. Straus and M. Hertzberg
214
a range of more- or less-specific agents targeting
different members of the HDAC family may
yield further information.
5.
6.
10.4
Conclusions
A variety of pharmacologic hypotheses have
been tested in the course of the last 50 years,
and none has been found entirely satisfactory
for predicting the outcomes of treatment. The
superiority of ABVD over MOPP is established.
Similarly, the more effective multiagent
BEACOPP regimen is being used in more and
more countries and groups. There appears to be
a potential trade-­off between the intensity of
chemotherapy and the value of consolidation
radiotherapy in advanced disease: it is not clear
whether any chemotherapy is intensive enough
for radiation to be dropped altogether, but functional imaging holds promise for lowering the
proportion of patients irradiated very
significantly.
As treatment has evolved, the balance between
toxicity and efficacy has been established, and
new approaches using response-adapted therapy
hold the promise of identifying the minority of
patients for whom early intensification is a necessity, while allowing de-escalation of treatment in
those destined to do well. The addition of brentuximab vedotin, the antibody-drug conjugate,
has slightly improved efficacy in the treatment of
stages III and IV HL. Finally, there are a small
number of novel agents currently undergoing
testing against recurrent and refractory disease
which appear to hold some promise.
References
1. Wilkinson JF, Fletcher F (1947) Effect of beta-­
chloroethylamine hydrochlorides in leukaemia,
Hodgkin’s disease, and polycythaemia vera; report
on 18 cases. Lancet 2(6476):540–545
2. Papac RJ (2001) Origins of cancer therapy. Yale J
Biol Med 74(6):391–398
3. Zubrod CG (1979) Historic milestones in curative
chemotherapy. Semin Oncol 6:490–505
4. Gross R, Lambers K (1958) Erste erfarhungen
in der Behandlung malignen tumoren mit einem
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
neuen N-lost phosphamidester. Dtsch Med Wschr
83:458–462
Rotolo V (1968) Vincaleukoblastine in the therapy
of malignant neoplasms. Friuli Med 23(1):31–52
Mathe G, Cattan A, Amiel JL, Schwarzenberg
L, Schneider M (1969) Experimental therapeutic
trials of leukemia and hematosarcomas: technologic and philosophic aspects. Ann N Y Acad Sci
164(3):776–792
Burchenal JH (1975) From wild fowl to stalking horses: alchemy in chemotherapy. Cancer
35:1121–1135
Gilman A (1963) The initial clinical trial of nitrogen
mustard. Am J Surg 105(574):578
Wagener DJT (2009) The history of oncology.
Springer, Berlin
Mathe G, Schweisguth O, Schneider M, Amiel JL,
Cattan A, Schwarzenberg L et al (1964) Value of
vincaleukoblastine in the treatment of Hodgkin’s
disease and other hematosarcomas and leukemias.
Sem Ther 40(5):320–324
Tubiana M, Henry-Amar M, Hayat M, Breur K,
Werf-Messing B, Burgers M (1979) Long-term
results of the E.O.R.T.C. randomized study of irradiation and vinblastine in clinical stages I and II of
Hodgkin’s disease. Eur J Cancer 15(5):645–657
Rosenberg SA, Kaplan HS (1966) Evidence for an
orderly progression in the spread of Hodgkin’s disease. Cancer Res 26(6):1225–1231
Tubiana M, Hayat M, Henry-Amar M, Breur K,
van der Werf MB, Burgers M (1981) Five-year
results of the E.O.R.T.C. randomized study of splenectomy and spleen irradiation in clinical stages I
and II of Hodgkin’s disease. Eur J Cancer 17(3):
355–363
Bergsagel DE, Alison RE, Bean HA, Brown TC,
Bush RS, Clark RM et al (1982) Results of treating
Hodgkin’s disease without a policy of laparotomy
staging. Cancer Treat Rep 66:717–731
Selby P, McElwain TJ, Canellos G (1987)
Chemotherapy for Hodgkin’s disease. Section I:
MOPP and its variants. In: Selby P, McElwain TJ
(eds) Hodgkin’s disease. Blackwell, Oxford
Younes A, Gopal AK, Smith SE, Ansell SM,
Rosenblatt JD, Savage KJ et al (2012) Results of a
pivotal phase II study of brentuximab vedotin for
patients with relapsed or refractory Hodgkin’s lymphoma. J Clin Oncol 30(18):2183–2189
Johnson PW, Federico M, Fossa A, Barrington SF,
Kirkwood A, Roberts TH et al (2013) Response
rates and toxicity of response-adapted therapy in
advanced Hodgkin lymphoma: initial results. From
The International RATHL Study. Hematologica
98(s2):2
Bernard J (1966) Current general principles of the
treatment of Hodgkin’s disease, lymphosarcoma and
reticulosarcoma. Rev Prat 16(7):871–879
Lacher MJ, Durant JR (1965) Combined vinblastine
and CHLORAMBUCIL therapy of HODGKIN’s
disease. Ann Intern Med 62:468–476
10 Principles of Chemotherapy in Hodgkin Lymphoma
20. DeVita VT Jr, Carbone PP (1967) Treatment of
Hodgkin’s disease. Med Ann 36(4):232–234
21. DeVita VT, Serpick AA, Carbone PP (1970)
Combination chemotherapy in the treatment of
advanced Hodgkin’s disease. Ann Intern Med
73(6):881–895
22. DeVita VT, Simon RM, Hubbard SM, Young RC,
Berard CW, Moxley JH et al (1980) Curability of
advanced Hodgkin’s disease with chemotherapy.
Ann Intern Med 92(5):587–595
23. Longo DL, Young RC, Wesley M, Hubbard SM,
Duffey PL, Jaffe ES et al (1986) Twenty years of
MOPP therapy for Hodgkin’s disease. J Clin Oncol
4(9):1295–1306
24. Canellos GP, Anderson JR, Propert KJ, Nissen
N, Cooper MR, Henderson ES et al (1992)
Chemotherapy of advanced Hodgkin’s disease with
MOPP, ABVD, or MOPP alternating with ABVD. N
Engl J Med 327(21):1478–1484
25. Carde P, MacKintosh FR, Rosenberg SA (1983) A
dose and time response analysis of the treatment of
Hodgkin’s disease with MOPP chemotherapy. J Clin
Oncol 1(2):146–153
26. Frei E III, Luce JK, Gamble JF, Coltman CA Jr,
Constanzi JJ, Talley RW et al (1973) Combination
chemotherapy in advanced Hodgkin’s disease:
induction and maintenance of remission. Ann Intern
Med 79(3):376–382
27. Somers R, Carde P, Henry-Amar M, Tarayre M,
Thomas J, Hagenbeek A et al (1994) A randomized
study in stage IIIB and IV Hodgkin’s disease comparing eight courses of MOPP versus an alteration
of MOPP with ABVD: a European Organization for
Research and Treatment of Cancer lymphoma cooperative group and Groupe Pierre-et-Marie-curie controlled clinical trial. J Clin Oncol 12(2):279–287
28. Carde P, Hayat M, Cosset JM, Somers R, Burgers
JM, Sizoo W et al (1988) Comparison of total nodal
irradiation versus combined sequence of mantle
irradiation with mechlorethamine, vincristine, procarbazine, and prednisone in clinical stages I and
II Hodgkin’s disease: experience of the European
Organization for Research and Treatment of Cancer.
NCI Monogr 6:303–310
29. Young RC, Chabner BA, Canellos GP, Schein PS,
DeVita VT (1973) Maintenance chemotherapy for
advanced Hodgkin’s disease in remission. Lancet
301(7816):1339–1343
30. Lowenbraun STAN, DeVita VT, Serpick AA (1970)
Combination chemotherapy with nitrogen mustard,
vincristine, Procarbazine and prednisone in previously treated patients with Hodgkin’s disease. Blood
36(6):704–717
31. Fisher RI, DeVita VT, Hubbard SP, Simon R, Young
RC (1979) Prolonged disease-free survival in
Hodgkin’s disease with MOPP reinduction after first
relapse. Ann Intern Med 90(5):761–763
32. Arseneau JC, Sponzo RW, Levin DL, Schnipper LE,
Bonner H, Young RC et al (1972) Non lymphomatous malignant tumors complicating Hodgkin’s dis-
215
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
ease : possible association with intensive therapy.
N Engl J Med 287(22):1119–1122
Weiden PL, Lerner KG, Gerdes A, Heywood JD,
Fefer A, Thomas ED (1973) Pancytopenia and leukemia in Hodgkin’s disease: report of three cases.
Blood 42(4):571–577
Sherins RJ, DeVita VT Jr (1973) Effect of drug treatment for lymphoma on male reproductive capacity.
Studies of men in remission after therapy. Ann Intern
Med 79(2):216–220
Corder MP, Young RC, Brown RS, Devita VT (1972)
Phytohemagglutinin-induced lymphocyte transformation: the relationship to prognosis of Hodgkin’s
disease. Blood 39(5):595–601
Crowther D, Wagstaff J, Deakin D, Todd I, Wilkinson
P, Anderson H et al (1984) A randomized study
comparing chemotherapy alone with chemotherapy
followed by radiotherapy in patients with pathologically staged IIIA Hodgkin’s disease. J Clin Oncol
2(8):892–897
Nicholson WM, Beard ME, Crowther D, Stansfeld
AG, Vartan CP, Malpas JS et al (1970) Combination
chemotherapy in generalized Hodgkin’s disease. Br
Med J 3(713):7–10
Ranson MR, Radford JA, Swindell R, Dikin DP,
Wilkinson PM, Harris M et al (1991) An analysis
of prognostic factors in stage III and IV Hodgkin’s
disease treated at a single centre with MVPP. Ann
Oncol 2(6):83–89
Dady PJ, McElwain TJ, Austin DE, Barrett A,
Peckham MJ (1982) Five years' experience with
ChlVPP: effective low-toxicity combination chemotherapy for hodgkin’s diseasechlvpp advanced
HL. Br J Cancer 45:851–859
McElwain TJ, Toy J, Smith E, Peckham MJ, Austin
DE (1977) A combination of chlorambucil, vinblastine, procarbazine and prednisolone for treatment of
Hodgkin’s disease. Br J Cancer 36(276):280
Luce JK, Gamble JF, Wilson HE, Monto RW, Isaacs
BL, Palmer RL et al (1971) Combined cyclophosphamide vincristine, and prednisone therapy of
malignant lymphoma. Cancer 28(2):306–317
Bonadonna G, Monfardini S (1969) Cardiac toxicity
of daunorubicin. Lancet 1(7599):837
Bonadonna G, Monfardini S, Oldini C (1969)
Comparative effects of vinblastine and procarbazine
in advanced Hodgkin’s disease. Eur J Cancer (1965)
5(4):393–402
Frei E, Luce JK, Talley RW, Veitkvicius VK, Wilson
HE (1972) 5-(3,3-dimethyl-1-triazeno)imidazole-­4-­
carboxamide (NSC-45388) in the treatment of lymphoma. Cancer Treat Rep 56(5):667–670
O'Bryan RM, Luce JK, Tailey RW, Gottlieb JA,
Baker LH, Bonadonna G (1973) Phase II evaluation
of adriamycin in human neoplasia. Cancer 32(1):1–8
Bonadonna G, Zucali R, Monfardini S, De Lena
M, Uslenghi C (1975) Combination chemotherapy
of Hodgkin’s disease with adriamycin, bleomycin, vinblastine, and imidazole carboxamide versus
MOPP. Cancer 36(1):252–259
D. Straus and M. Hertzberg
216
47. Duggan DB, Petroni GR, Johnson JL, Glick JH,
Fisher RI, Connors JM et al (2003) Randomized
comparison of ABVD and MOPP/ABV hybrid for
the treatment of advanced Hodgkin’s disease: report
of an intergroup trial. J Clin Oncol 21(4):607–614
48. Boleti E, Mead GM (2007) ABVD for Hodgkin’s
lymphoma: full-dose chemotherapy without
dose reductions or growth factors. Ann Oncol
18(2):376–380
49. Martin WG, Ristow KM, Habermann TM, Colgan
JP, Witzig TE, Ansell SM (2005) Bleomycin pulmonary toxicity has a negative impact on the outcome
of patients with Hodgkin’s lymphoma. J Clin Oncol
23(30):7614–7620
50. Straus DJ, Portlock CS, Qin J, Myers J, Zelenetz
AD, Moskowitz C et al (2004) Results of a prospective randomized clinical trial of doxorubicin,
bleomycin, vinblastine, and dacarbazine (ABVD)
followed by radiation therapy (RT) versus ABVD
alone for stages I, II, and IIIA nonbulky Hodgkin
disease. Blood 104(12):3483–3489
51. Canellos GP, Duggan D, Johnson J, Niedzwiecki D
(2004) How important is bleomycin in the adriamycin + bleomycin + vinblastine + dacarbazine regimen? J Clin Oncol 22(8):1532–1533
52. Johnson P Adapted treatment guided by interim
PET-CT scan in advanced Hodgkin’s lymphoma. N
Engl J Med 374(25):2419–2429
53. Norton L, Simon R (1977) Tumor size, sensitivity to
therapy, and design of treatment schedules. Cancer
Treat Rep 61:1307–1317
54. Diehl V, Franklin J, Pfreundschuh M, Lathan B,
Paulus U, Hasenclever D et al (2003) Standard and
increased-dose BEACOPP chemotherapy compared
with COPP-ABVD for advanced Hodgkin’s disease.
N Engl J Med 348(24):2386–2395
55. Linch DC, Winfield D, Goldstone AH, Moir D,
Hancock B, McMillan A et al (1993) Dose intensification with autologous bone-marrow transplantation in relapsed and resistant Hodgkin’s
disease: results of a BNLI randomised trial. Lancet
341(8852):1051–1054
56. Schmitz N, Pfistner B, Sextro M, Sieber M, Carella
AM, Haenel M et al (2002) Aggressive conventional chemotherapy compared with high-dose
chemotherapy with autologous haemopoietic
stem-cell transplantation for relapsed chemosensitive Hodgkin’s disease: a randomised trial. Lancet
359(9323):2065–2071
57. Hasenclever D, Brosteanu O, Gerike T, Loeffler M
(2001) Modelling of chemotherapy: the effective
dose approach. Ann Hematol 80(Suppl 3):B89–B94
58. Hasenclever D, Loeffler M, Diehl V (1996) Rationale
for dose escalation of first line conventional chemotherapy in advanced Hodgkin’s disease. German
Hodgkin’s Lymphoma Study Group. Ann Oncol
7(Suppl 4):95–98
59. Owadally WS, Sydes MR, Radford JA, Hancock
BW, Cullen MH, Stenning SP et al (2010) Initial
dose intensity has limited impact on the out-
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
come of ABVD chemotherapy for advanced
Hodgkin lymphoma (HL): data from UKLG LY09
(ISRCTN97144519). Ann Oncol 21:568–573
Brosteanu O, Hasenclever D, Loeffler M, Diehl V,
Group GH (2004) Low acute hematological toxicity
during chemotherapy predicts reduced disease control in advanced Hodgkin’s disease. Ann Hematol
83(3):176–182
Ribrag V, Koscielny S, Casasnovas O, Cazeneuve
C, Brice P, Morschhauser F et al (2009)
Pharmacogenetic study in Hodgkin lymphomas
reveals the impact of UGT1A1 polymorphisms on
patient prognosis. Blood 113(14):3307–3313
Engert A, Diehl V, Franklin J, Lohri A, Dorken B,
Ludwig WD et al (2009) Escalated-dose BEACOPP
in the treatment of patients with advanced-stage
Hodgkin’s lymphoma: 10 years of follow-up
of the GHSG HD9 study. J Clin Oncol 27(27):
4548–4554
Henry-Amar M, Hayat M, Meerwaldt JH, Burgers
M, Carde P, Somers R et al (1990) Causes of death
after therapy for early stage Hodgkin’s disease
entered on EORTC protocols. EORTC lymphoma
cooperative group. Int J Radiat Oncol Biol Phys
19(5):1155–1157
Van Leeuwen FE, Klokman WJ, Hagenbeek A,
Noyon R, Van den Beltdusebout AW, Van Kerkhoff
EHM et al (1994) 2nd cancer risk following
hodgkins-­disease - a 20-year follow-up-study. J Clin
Oncol 12(2):312–325
Bartlett NL, Rosenberg SA, Hoppe RT, Hancock SL,
Horning SJ (1995) Brief chemotherapy, Stanford V,
and adjuvant radiotherapy for bulky or advanced-­
stage Hodgkin’s disease: a preliminary report. J Clin
Oncol 13(5):1080–1088
Horning SJ, Williams J, Bartlett NL, Bennett JM,
Hoppe RT, Neuberg D et al (2000) Assessment of the
Stanford V regimen and consolidative radiotherapy
for bulky and advanced Hodgkinís disease: eastern
cooperative oncology group pilot study E1492. J
Clin Oncol 18(5):972
Horning SJ, Hoppe RT, Breslin S, Bartlett NL,
Brown BW, Rosenberg SA (2002) Stanford V and
radiotherapy for locally extensive and advanced
Hodgkin’s disease: mature results of a prospective
clinical trial. J Clin Oncol 20(3):630–637
Gobbi PG, Levis A, Chisesi T, Broglia C, Vitolo
U, Stelitano C et al (2005) ABVD versus modified
Stanford V versus MOPPEBVCAD with optional
and limited radiotherapy in intermediate- and
advanced-stage Hodgkin’s lymphoma: final results
of a multicenter randomized trial by the Intergruppo
Italiano Linfomi. J Clin Oncol 23(36):9198–9207
Hoskin PJ, Lowry L, Horwich A, Jack A, Mead B,
Hancock BW et al (2009) Randomized comparison
of the Stanford V regimen and ABVD in the treatment of advanced Hodgkin’s lymphoma: United
Kingdom National Cancer Research Institute lymphoma group study ISRCTN 64141244. J Clin
Oncol 27(32):5390–5396
10 Principles of Chemotherapy in Hodgkin Lymphoma
70. Gordon LI, Hong F, Fisher RI, Bartlett NL, Connors
JM, Gascoyne RD et al (2013) Randomized phase
III trial of ABVD versus Stanford V with or without
radiation therapy in locally extensive and advanced-­
stage Hodgkin lymphoma: an intergroup study coordinated by the eastern cooperative oncology group
(E2496). J Clin Oncol 31(6):684–691
71. Radford JA, Rohatiner AZS, Ryder WDJ, Deakin
DP, Barbui T, Lucie NP et al (2002) ChlVPP/EVA
hybrid versus the weekly VAPEC-B regimen for previously untreated Hodgkin’s disease. J Clin Oncol
20(13):2988–2994
72. Green JA, Dawson AA, Fell LF, Murray S (1980)
Measurement of drug dosage intensity in MVPP
therapy in Hodgkin’s disease. Br J Clin Pharmacol
9:511–514
73. Diehl V, Sieber M, Ruffer U, Lathan B, Hasenclever
D, Pfreundschuh M et al (1997) BEACOPP: an intensified chemotherapy regimen in advanced Hodgkin’s
disease. The German Hodgkin’s lymphoma study
group. Ann Oncol 8(2):143–148
74. Diehl V, Haverkamp H, Mueller R, Mueller-­
Hermelink H, Cerny T, Markova J et al (2009) Eight
cycles of BEACOPP escalated compared with 4
cycles of BEACOPP escalated followed by 4 cycles
of BEACOPP baseline with or without radiotherapy
in patients in advanced stage Hodgkin lymphoma
(HL): final analysis of the HD12 trial of the German
Hodgkin study group (GHSG). ASCO Meet Abst
27(15S):8544
75. Engert A, Franklin J, Mueller RP, Eich HT,
Gossmann A, Mueller-Hermelink HK et al (2006)
HD12 randomised trial comparing 8 dose-escalated
cycles of BEACOPP with 4 escalated and 4 baseline
cycles in patients with advanced stage Hodgkin lymphoma (HL): an analysis of the German Hodgkin
lymphoma study group (GHSG), University of
Cologne, Cologne, Germany. ASH Annu Meet Abstr
108(11):99
76. Kobe C, Dietlein M, Franklin J, Markova J, Lohri A,
Amthauer H et al (2008) Positron emission tomography has a high negative predictive value for progression or early relapse for patients with residual disease
after first-line chemotherapy in advanced-stage
Hodgkin lymphoma. Blood 112(10):3989–3994
77. Federico M, Luminari S, Iannitto E, Polimeno G,
Marcheselli L, Montanini A et al (2009) ABVD
compared with BEACOPP compared with CEC
for the initial treatment of patients with advanced
Hodgkin’s lymphoma: results from the HD2000
Gruppo Italiano per lo studio Dei Linfomi trial. J
Clin Oncol 27(5):805–811
78. Gianni AM, Rambaldi A, Zinzani P, Levis A,
Brusamolino E, Pulsoni A et al (2008) Comparable
3-year outcome following ABVD or BEACOPP
first-line chemotherapy, plus pre-planned high-dose
salvage, in advanced Hodgkin lymphoma (HL): a
randomized trial of the Michelangelo, GITIL and
IIL cooperative groups. ASCO Meet Abstr 26(15
Suppl):8506
217
79. Viviani S, Zinzani PL, Rambaldi A, Brusamolino
E, Levis A, Bonfante V et al (2011) ABVD versus BEACOPP for Hodgkin’s lymphoma when
high-dose salvage is planned. N Engl J Med
365(3):203–212
80. Carde P, Karrasch M, Fortpied C, Brice P, Khaled
HM, Caillot D et al (2012) ABVD (8cycles) versus
BEACOPP (4 escalated cycles => 4 baseline) in
stage III-IV high-risk Hodgkin Lymphoma (HL):
First results of EORTC 20012 Intergroup randomized phase III clinical trial. J Clin Oncol 30:abstr
8002
81. von Tresckow B, Haverkamp H, Boll B, Eichenauer
DA, Sasse S, Fuchs M et al (2013) Impact of dose
reduction of bleomycin and vincristine in patients
with advanced Hodgkin lymphoma treated with
BEACOPP: A comprehensive analysis of the
German Hodgkin Study Group (GHSG) HD12 and
HD15 trials. Blood 122(21):Abstract 637
82. Proctor SJ, Mackie M, Dawson A, Prescott B,
Lucraft HL, Angus B et al (2002) A population-­
based study of intensive multi-agent chemotherapy with or without autotransplant for the highest
risk Hodgkin’s disease patients identified by the
Scotland and Newcastle lymphoma group (SNLG)
prognostic index: a Scotland and Newcastle lymphoma group study (SNLG HD III). Eur J Cancer
38(6):795–806
83. Federico M, Bellei M, Brice P, Brugiatelli M, Nagler
A, Gisselbrecht C et al (2003) High-dose therapy and
autologous stem-cell transplantation versus conventional therapy for patients with advanced Hodgkin’s
lymphoma responding to front-line therapy. J Clin
Oncol 21(12):2320–2325
84. Arakelyan N, Berthou C, Desablens B, De Guibert
S, Delwail V, Moles MP et al (2008) Radiation
therapy versus 4 cycles of combined doxorubicin, bleomycin, vinblastine, and Dacarbazine plus
Myeloablative chemotherapy with autologous stem
cell transplantation five-year results of a randomized trial on behalf of the GOELAMS group. Cancer
113(12):3323–3330
85. Cheson BD, Fisher RI, Barrington SF,
Cavalli F, Schwartz LH, Zucca E et al (2014)
Recommendations for initial evaluation, staging, and
response assessment of Hodgkin and non-Hodgkin
lymphoma: the Lugano classification. J Clin Oncol
32(27):3059–3068
86. Hutchings M, Kostakoglu L, Zaucha JM, Malkowski
B, Biggi A, Danielewicz I et al (2014) In vivo treatment sensitivity testing with positron emission
tomography/computed tomography after one cycle
of chemotherapy for Hodgkin lymphoma. J Clin
Oncol 32(25):2705–2711
87. Straus DJ, Johnson JL, LaCasce AS, Bartlett NL,
Kostakoglu L, Hsi ED et al (2011) Doxorubicin,
vinblastine, and gemcitabine (CALGB 50203) for
stage I/II nonbulky Hodgkin lymphoma: pretreatment prognostic factors and interim PET. Blood
117(20):5314–5320
D. Straus and M. Hertzberg
218
88. Radford J, Illidge T, Counsell N, Hancock B,
Pettengell R, Johnson P et al (2015) Results of a trial
of PET-directed therapy for early-stage Hodgkin’s
lymphoma. N Engl J Med 372(17):1598–1607
89. Straus D, Pitcher B, Kostakoglu L, Grecula JC, Hsi
ED, Schoder H et al (2016) Results of US intergroup trial of response-adapted chemotherapy or
chemotherapy/radiation therapy based on PET
for ­non-­bulky stage I and II Hodgkin lymphoma
(CALGB/ALLIANCE 50604) (Abstract). Haematol
J Eur Hematol Assoc 101:13. ISHL 10. Journal of the
European Hematology Association. 101. Cologne,
Germany: Ferrata Storti Foundation
90. Raemaekers JM, Andre MP, Federico M, Girinsky T,
Oumedaly R, Brusamolino E et al (2014) Omitting
radiotherapy in early positron emission tomography-­
negative stage I/II Hodgkin lymphoma is associated with an increased risk of early relapse: clinical
results of the preplanned interim analysis of the randomized EORTC/LYSA/FIL H10 trial. J Clin Oncol
32(12):1188–1194
91. Andre MPE, Girinsky T, Federico M, Reman O,
Fortpied C, Gotti M et al (2017) Early positron
emission tomography response-adapted treatment in
stage I and II Hodgkin lymphoma: final results of
the randomized EORTC/LYSA/FIL H10 trial. J Clin
Oncol 35(16):1786–1794
92. Fuchs M, Goergen H, Kobe C, Eich H, Baues C,
Greil R et al (2018) PET-guided treatment of early-­
stage favorable Hodgkin lymphoma: final results of
the international, randomized phase 3 trial HD16 by
the German Hodgkin Study Group. Blood 132(Suppl
1):925
93. Press OW, Li H, Schoder H, Straus DJ, Moskowitz
CH, LeBlanc M et al (2016) US intergroup trial of
response-adapted therapy for stage III to IV Hodgkin
lymphoma using early interim fluorodeoxyglucose-­
positron emission tomography imaging: southwest oncology group S0816. J Clin Oncol
34(17):2020–2027
94. Borchmann P, Goergen H, Kobe C, Lohri A, Greil R,
Eichenauer DA et al (2018) PET-guided treatment in
patients with advanced-stage Hodgkin’s lymphoma
(HD18): final results of an open-label, international,
randomised phase 3 trial by the German Hodgkin
Study Group. Lancet 390(10114):2790–2802
95. Connors JM, Jurczak W, Straus DJ, Ansell SM, Kim
WS, Gallamini A et al (2018) Brentuximab vedotin
with chemotherapy for stage III or IV Hodgkin’s
lymphoma. N Engl J Med 378(4):331–344
96. Ansell SM, Horwitz SM, Engert A, Khan KD, Lin T,
Strair R et al (2007) Phase I/II study of an anti-CD30
monoclonal antibody (MDX-060) in Hodgkin’s lymphoma and anaplastic large-cell lymphoma. J Clin
Oncol 25(19):2764–2769
97. Younes A, Romaguera J, Hagemeister F, McLaughlin
P, Rodriguez MA, Fiumara P et al (2003) A pilot
study of rituximab in patients with recurrent, classic
Hodgkin disease. Cancer 98(2):310–314
98. Ekstrand BC, Lucas JB, Horwitz SM, Fan Z, Breslin
S, Hoppe RT et al (2003) Rituximab in lymphocyte-­
99.
100.
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
predominant Hodgkin disease: results of a phase 2
trial. Blood 101(11):4285–4289
Mendler JH, Kelly J, Voci S, Marquis D, Rich L,
Rossi RM et al (2008) Bortezomib and gemcitabine
in relapsed or refractory Hodgkin’s lymphoma. Ann
Oncol 19(10):1759–1764
Younes A, Pro B, Fanale M, McLaughlin P, Neelapu
S, Fayad L et al (2007) Isotype-selective HDAC
inhibitor MGCD0103 decreases serum TARC
concentrations and produces clinical responses in
heavily pretreated patients with relapsed classical
Hodgkin lymphoma. Blood 110(11):2566
Bonadonna G, Santoro A, Bonfante V, Valagussa P
(1982) Cyclic delivery of MOPP and ABVD combinations in stage IV Hodgkin’s disease: rationale,
background studies, and recent results. Cancer Treat
Rep 66(4):881–887
Radford JA, Crowther D, Rohatiner AZ, Ryder
WD, Gupta RK, Oza A et al (1995) Results of a
randomized trial comparing MVPP chemotherapy
with a hybrid regimen, ChlVPP/EVA, in the initial treatment of Hodgkin’s disease. J Clin Oncol
13(9):2379–2385
Sutcliffe S, Wrigley PFM, Peto J, Lister TA,
Stansfeld AG, Whitehouse JM et al (1978) MVPP
chemotherapy regimen for advanced Hodgkin’s disease. Br Med J 6114:679–683
Selby P, Patel P, Milan S, Meldrum M, Mansi J,
Mbidde E et al (1990) ChlVPP combination chemotherapy for Hodgkin’s disease: long-term results. Br
J Cancer 62(2):279–285
Johnson PWM, Radford JA, Cullen MH, Sydes
MR, Walewski J, Jack AS et al (2005) Comparison
of ABVD and alternating or hybrid multidrug regimens for the treatment of advanced Hodgkin’s lymphoma: results of the United Kingdom lymphoma
group LY09 trial (ISRCTN97144519). J Clin Oncol
23(36):9208–9218
Santoro A, Bonadonna G, Bonfante V, Valagussa P
(1982) Alternating drug combinations in the treatment of advanced Hodgkin’s disease. N Engl J Med
306(13):770–775
Viviani S, Bonadonna G, Santoro A, Bonfante V,
Zanini M, Devizzi L et al (1996) Alternating versus
hybrid MOPP and ABVD combinations in advanced
Hodgkin’s disease: ten-year results. J Clin Oncol
14(5):1421–1430
Sieber M, Tesch H, Pfistner B, Rueffer U, Paulus
U, Munker R et al (2004) Treatment of advanced
Hodgkin’s disease with COPP/ABV/IMEP versus
COPP/ABVD and consolidating radiotherapy: final
results of the German Hodgkin’s lymphoma study
group HD6 trial. Ann Oncol 15(2):276–282
Aleman BM, Raemaekers JM, Tirelli U, Bortolus R
(2003) T veer MB, Lybeert ML, et al. involved-field
radiotherapy for advanced Hodgkin’s lymphoma. N
Engl J Med 348(24):2396–2406
Klimo P, Connors JM (1985) MOPP/ABV hybrid
program: combination chemotherapy based on early
introduction of seven effective drugs for advanced
Hodgkin’s disease. J Clin Oncol 3(9):1174–1182
Treatment of Early Favorable
Hodgkin Lymphoma
11
Wouter Plattel and Pieternella Lugtenburg
Contents
11.1
Introduction 219
11.2
11.2.1
11.2.2
efining Favorable Early-­Stage Disease D
Staging Prognostic Factors 220
220
221
11.3
Radiotherapy Alone 222
11.4
Late Treatment Effects and Mortality 223
11.5
11.5.1
11.5.2
11.5.3
11.5.4
11.5.5
Combined Modality Treatment Radiotherapy Alone Versus CMT Optimal Number of Cycles of Chemotherapy Optimal Chemotherapy Combination Optimal Radiation Dose Optimal Radiation Field Size 224
225
227
227
227
228
11.6
Chemotherapy Alone 230
11.7
Treatment Adaptation Based on PET Scan Response 230
11.8
Recommendations and Future Directions 234
References 234
11.1
W. Plattel (*)
Department of Haematology, University Medical
Center Groningen, University of Groningen,
Groningen, The Netherlands
e-mail: w.j.plattel@umcg.nl
P. Lugtenburg
Department of Haematology, Erasmus MC Cancer
Institute, University Medical Center,
Rotterdam, The Netherlands
Introduction
Historically, Hodgkin lymphoma (HL) was the
first malignant disease that could be cured. In the
past century, the first successful outcomes of
radiotherapy employing large radiation fields
were reported, in particular, in patients with limited disease.
Further refinement of this initial treatment
approach was achieved through carefully
designed prospective randomized phase III
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_11
219
W. Plattel and P. Lugtenburg
220
c­linical trials. In this context, the step-by-step
development of uniformly accepted staging procedures and clear definitions of stages and
response criteria was a major achievement. This
allowed direct comparison of study results performed in different consortia worldwide.
Focusing on stage-adapted treatment of HL,
these trials allowed the definition of clinical prognostic factors. These, in turn, lead to risk-­adapted
treatment, which became more refined with subsequent studies. In line with these advances, treatment strategies changed from radiotherapy only
using extended-field radiotherapy (EFRT) and
later involved-field radiotherapy (IFRT) to combined modality treatment (CMT) with rather small
radiotherapy fields and chemotherapy exposure.
Thanks to the long-term follow-up of thousands of patients treated within clinical trials over
decades, significant late effects of treatment
became apparent. The higher mortality rate in HL
survivors turned out to be mainly due to secondary malignancies and damage to the cardiovascular and respiratory systems. Based on these
unexpected findings, the ingredients of curative
regimens were further adjusted. As far as possible, noncarcinogenic cytostatic agents were
introduced in newly developed chemotherapy
regimens, and radiation doses were further
reduced. This has led to the current major challenges in the treatment of early-stage HL: maintaining the very high cure rates and at the same
time reducing the incidence of early and late toxicity. To further improve on this strategy, it is
strongly advocated to treat early-stage HL
patients within clinical trials.
This chapter deals with past and recent developments in the treatment of stage I and II HL
with favorable prognostic factors comprising
about 40% of all early-stage HL patients.
11.2
efining Favorable Early-­
D
Stage Disease
11.2.1 Staging
In HL patients, prognosis is distinctly worse with
each progressive stage of disease, and the selec-
tion of appropriate treatment depends on accurate
staging of the extent of disease. The Ann Arbor
staging classification was formulated in 1971 and
is still the most commonly used staging system
for HL [1]. During the Cotswold meeting in
1989, some modifications were introduced to
account for new imaging techniques such as
computerized tomography (CT) scanning. In
addition, clinical involvement of the liver and
spleen was redefined, to formally introduce the
concept of bulky disease and to draw the attention to the problem of equivocal complete remission [2]. Stage I indicates involvement of a single
lymph node region or a single extranodal organ
or site. In stage II disease, two or more lymph
node regions on the same side of the diaphragm
are involved, or there is localized involvement of
an extranodal organ or site and of one or more
lymph node regions on the same side of the diaphragm. The stage number is followed by the suffix A or B indicating the absence (A) or presence
(B) of one or more of the following constitutional
symptoms: (a) unexplained fever with temperatures above 38 °C during the previous month, (b)
drenching night sweats during the previous
months, and (c) unexplained weight loss of more
than 10% of body weight in the previous
6 months. Mediastinal bulk was defined by the
ratio of the maximum transverse tumor diameter
to the internal thoracic diameter at the level of the
T5–T6 vertebral interspace. A ratio exceeding
one-third was considered bulky.
For the initial staging of HL, a detailed history, complete physical examination, and imaging studies with whole body positron emission
tomography using [18F]-fluoro-2-deoxy-dglucose (FDG-PET, here referred to as PET)
scanning and CT scans of the neck, thorax, abdomen, and pelvis are generally recommended [3,
4]. In patients with PET-CT-assessed HL, bone
marrow biopsy can be omitted [5]. See Chaps. 6
and 7 for a more comprehensive review of clinical evaluation and functional imaging.
About 8% of stage I–II HL patients present
with infradiaphragmatic disease [6, 7]. Patients
with infradiaphragmatic HL are generally older,
more frequently male, have poorer performance
status, and present less frequently with nodular
11
Treatment of Early Favorable Hodgkin Lymphoma
221
Table 11.1 Definition of early-stage favorable HL
EORTC–GELA
CS I–II without risk factors
(supradiaphragmatic):
– No large mediastinal mass
– Age <50 years
– No elevated ESRa
– 1–3 involved nodal regions
GHSG
CS I–II without risk
factors:
– No large mediastinal
mass
– No extranodal disease
– No elevated ESRa
– 1–2 involved nodal
regions
NCI-C/ECOG
CS I–IIA without risk factors
(supradiaphragmatic):
– No large mediastinal mass
– Age <40 years
– ESR <50 mm/h
– 1–3 involved nodal regions
– LPHL or NS histology
EORTC European Organization for Research and Treatment of Cancer, GELA Groupe d’Etude des Lymphomes de
l’Adulte, GHSG German Hodgkin Study Group, NCI-C National Cancer Institute of Canada, ECOG Eastern Cooperative
Oncology Group, CS clinical stage, ESR erythrocyte sedimentation rate, LPHL nodular lymphocyte-predominant
Hodgkin lymphoma, NS nodular sclerosis
a
ESR <50 mm/h without B symptoms or ESR <30 mm/h with B symptoms
sclerosis subtype compared to patients with
supradiaphragmatic disease. Furthermore, these
patients have a significantly poorer progression-­
free survival (PFS) and overall survival (OS) as
compared to patients with supradiaphragmatic
disease [7]. Therefore, these patients should be
considered as early unfavorable HL which is further described in Chap. 12.
11.2.2 Prognostic Factors
Historically, several studies describing prognostic factors in early-stage HL have been performed
[8, 9] to predict for occult disease in the abdomen
and effectiveness of treatment. They were derived
from long-term follow-up of patient cohorts
treated in a variety of phase III prospective randomized trials. The prognostic significance of
bulky disease particularly in the mediastinum,
the presence of constitutional symptoms, the
erythrocyte sedimentation rate (ESR), and the
number of involved lymph node regions were
uniformly included in clinically applied prognostic models (see Chap. 8 for prognostic factors).
Different Lymphoma Collaborative Groups
worldwide use varying combinations of prognostic factors to identify prognostic risk groups.
These prognostic factors allow patients to be
stratified into favorable or unfavorable prognostic groups. The current definitions of a favorable
treatment group according to the different study
groups in Europe and the United States are presented in Table 11.1. The Lymphoma Group of
the European Organization for Research and
Treatment of Cancer (EORTC) and the French–
Belgian Groupe d’Etude des Lymphomes de
l’Adulte (GELA) define clinical stage I–II
patients as favorable if they present with the following characteristics: age <50 years and low
ESR (<50 mm/h without and <30 mm/h with B
symptoms), no more than three involved lymph
node regions, and no large mediastinal mass [10].
All these criteria need to be met to be “favorable.” The German Hodgkin Study Group
(GHSG) criteria differ slightly in that they substituted age <50 years with no extranodal disease
and specify no more than two involved nodal
regions rather than ≤3 as in the EORTC [11]. In
Canada and North America, it is common to
define an early or limited stage risk group as
stage I and IIA disease without bulky disease (see
Table 11.1).
Many of these defined risk factors are reflective of or correlate with disease burden. Currently
applied PET/CT imaging for staging of HL
allows accurate measurement of total metabolic
tumor volume (TMTV). The prognostic value of
TMTV has been increasingly described in HL. In
early-stage HL, a retrospective analyses of
TMTV of staging PET/CT images from the
EORTC/GELA/FIL H10 showed that TMTV
outperforms the classical risk factors described
above in terms of prediction of interim PET
W. Plattel and P. Lugtenburg
222
p­ ositivity and PFS [12]. However, standardization of measurement of TMTV is the major challenge before clinical application.
11.3
Radiotherapy Alone
The use of radiation therapy, pioneered at
Stanford University in the 1960s by Henry
Kaplan and Saul Rosenberg, offered HL patients
the first hope for cure. In the treatment of early
stages, EFRT was considered the standard treatment modality for many years. With this technique, radiation was delivered not only to the
clinically involved but also to the adjacent, clinically uninvolved sites. Because it was known that
HL spreads to contiguous nodal sites, mantle
field RT encompassed all nodal sites above the
diaphragm. The combination of mantle field with
inverted-Y field and spleen irradiation was
termed “subtotal nodal irradiation” (STNI). See
Chap. 9 for definitions of field size.
Significant advances in the treatment of HL
were then derived from clinical trials.
Investigators at Stanford demonstrated that
­radiation therapy alone using total lymphoid irradiation or STNI is an adequate treatment for
nearly all patients with pathologic stages I–II. In
a series of 109 patients, the freedom from relapse
rate at 10 years was 77% [13].
A retrospective study from Canada studied
the impact of patient selection and EFRT on
outcome among patients with clinical stages I
and II treated between 1978 and 1986. Patients
with favorable prognostic features (age
<50 years, ESR <40 mm/h, and lymphocytepredominant or nodular sclerosing histology)
treated with mantle and para-aortic-splenic irradiation had only 12.7% actuarial risk of relapse
at 8 years [14].
Between 1964 and 1987, the EORTC performed four consecutive randomized clinical trials aiming to delineate the subsets of patients
with stage I and II disease who could be safely
treated with RT alone [15, 16] (Table 11.2).
Taken together, these four randomized trials
demonstrated that staging laparotomy could be
safely omitted in patients with favorable clinical
characteristics in early favorable HL and that
these patients could be treated by STNI (40 Gy)
with a similar outcome as obtained by staging
laparotomy followed by mantle field RT (40 Gy).
Another important finding was that the overall
outcome had gradually improved over the years
(Fig. 11.1).
The total radiation dose in these EORTC trials
was always 40 Gy. The GHSG HD4 trial showed
that patients without risk factors had similar outcomes when treated with 40 Gy radiaton to the
involved field and 30 Gy to the non-involved
extended field [22]. The 7-year relapse-free and
overall survival rates were 78% vs. 83% and 91%
vs. 96%, respectively.
Radiation in mantle field technique was
expected to cause less long-term toxicity compared with STNI. However, in clinically staged
patients, results with mantle field irradiation
alone have been disappointing. In the EORTC
H7-VF and H8-VF trials, 40 female patients were
treated with mantle field RT only. The respective
prognostic factors were stage IA, age <40 years,
nodular sclerosing or lymphocyte-predominant
histology, and ESR <50 mm/h. These patients
were expected to have a very low risk of occult
abdominal involvement (5%). The relapse-free
survival was however lower than expected: a total
of 23% had relapsed at 6 years [21]. Because of
this unacceptable rate, the very favorable subgroup has since been treated according to the
EORTC strategy for the favorable subgroup.
Specht et al. reported on the influence of
radiation field size on long-term outcome in
early-­stage disease in a meta-analysis of eight
randomized trials evaluating larger vs. smaller
radiation fields [23]. These trials included
almost 2000 patients with both, favorable and
unfavorable prognosis stage I–II disease. A
definite and substantial reduction in the risk of
treatment failure was demonstrated if more
extensive radiotherapy was used. The 10-year
risk of recurrence was 43% for patients treated
with smaller-field irradiation compared to 31%
for those treated with larger-field radiation
therapy. Although the additional radiotherapy
prevented a substantial proportion of recurrences, it did not significantly affect overall
11
Treatment of Early Favorable Hodgkin Lymphoma
223
Table 11.2 Early-stage favorable HL: selection of randomized studies of radiotherapy alone
Number of
patients
Outcome
288
A. 38% DFS
(15 years)
B. 60% DFS
(15 years)
p < 0.001
Trial
EORTC HI
Year
1964–1971
Study arms
A. M
antle field or
inverted-Y RT
B. The same RT followed
by vinblastine
EORTC H2
1972–1976
A. Laparotomy and mantle
field + Para-aortic
lymph node RT
B. STNI
300
Laparotomy negative
patients
A. Mantle field RT
198
EORTC
H5F
1977–1982
B. STNI
EORTC
H6F
EORTC
H7VF-­
H8VF
GHSG
HD4
1982–1987
A. Laparotomy, if negative:
Mantle field RT for LP
or NSc histology
STNI for MC or LD
histology
B. STNI
262
1988–1993
Mantle field RT
40
1988–1994
A. STNI 40 Gy
376
B. STNI 30 Gy + IFRT
10 Gy
A. 76% DFS
(12 years)
Overall
survival
A. 58% OS
(15 years)
B. 65% OS
(15 years)
p = 0.15
(NS)
A. 79% OS
(12 years)
B. 68% DFS
(12 years)
p = 0.18
(NS)
A. 69% DFS
(9 years)
B. 70% DFS
(9 years)
p > 0.50
(NS)
A. 84% RFS
(6 years)
B. 77% OS
(12 years)
p = 0.38
(NS)
A. 94% OS
(9 years)
B. 91% OS
(9 years)
p > 0.50
(NS)
A. 89% OS
(6 years)
Carde et al.
[20]
B. 80% RFS
(6 years)
p = 0.25
(NS)
RFS 73%
(6 years)
B. 93% OS
(6 years)
p = 0.24
(NS)
OS 95%
(6 years)
Noordijk
et al. [21]
A. 78% RFS
(7 years)
B. 83% RFS
(7 years)
p = 0.093
(NS)
A. 91% OS
(7 years)
B. 96% OS
(7 years)
p = 0.16
(NS)
Reference
Tubiana
et al. [17]
Tubiana
et al.
[16, 18]
Carde et al.
[19]
Dühmke
et al. [22]
EORTC European Organization for Research and Treatment of Cancer, GHSG German Hodgkin Study Group, DFS
disease-free survival, OS overall survival, RFS relapse-free survival, STNI subtotal nodal irradiation, RT radiotherapy,
IFRT involved-field radiotherapy, Gy Gray, NS not significant, LP lymphocyte predominant, NSc nodular sclerosing,
MC mixed cellularity, LD lymphocyte depleted
mortality. The lack of survival difference suggests that salvage chemotherapy for relapse
after initial radiotherapy is effective enough to
minimize the impact of any increase in relapse
on survival.
To summarize, STNI was considered a standard treatment for early favorable HL until the
1990s. However, 25–30% of patients eventually
relapsed with subsequent 10-year survival rates
of only 63% [24].
11.4
Late Treatment Effects
and Mortality
As the number of patients surviving HL increased
and there was longer follow-up, it became evident that their life expectancy did not revert to
that of the age-matched general population. The
higher mortality of HL patients is largely a result
of the long-term effects of treatment. Important
late effects comprise secondary malignancies,
W. Plattel and P. Lugtenburg
224
Fig. 11.1 Disease-free
survival and overall
survival in consecutive
EORTC Lymphoma
Group trials on
early-stage favorable
Hodgkin lymphoma
(HL). DFS disease-free
survival, OS overall
survival
100
80
60
40
20
0
H1
H2
DFS
H5F
H6F
H7F
H8F
OS
cardiovascular diseases, pulmonary problems,
gonadal dysfunction, infectious complications,
and fatigue. The incidence of the most life-­
threatening late side effects, i.e., secondary cancers and cardiovascular diseases, is significantly
related to the radiation dose and field size, choice
of cytostatic drugs, and total amount of drugs
administered.
In patients with early favorable disease, mortality from causes other than HL has increased
over time, exceeding HL-related mortality after
10–15 years [25, 26]. A large study with a median
follow-up of more than 17 years examined case-­
specific mortality and absolute excess mortality,
compared to population rates, in a cohort of 1261
Dutch patients [25]. These patients were younger
than 40 years when treated between 1965 and
1987. HL was the most frequent cause of death
(55%), followed by secondary malignancies
(22%) and cardiovascular diseases (9%). In the
first 10 years following initial treatment, the
excess mortality rate was largely due to the primary disease, while after 10 years, causes other
than HL contributed most to excess mortality.
The actuarial risk of death is shown in Fig. 11.2.
Even after 30 years of follow-up, there was no
evidence of a decline in the relative risk of death
from causes other than HL. In 30-year survivors,
the annual excess mortality rate from all causes
other than HL was nearly 3 per 100 patients.
Solid tumors, especially in the digestive and
respiratory tract, contributed most to this excess
risk, followed by cardiovascular diseases [25]. In
2009, the EORTC and the GELA published their
results of a study analyzing the cause-specific
excess mortality in adult patients with respect to
treatment modality [27]. The study population
consisted of 4401 patients aged 15–69 in all
stages, who were treated between 1964 and 2000.
In patients with early-stage disease, the overall
excess mortality was associated with age
≥40 years (p = 0.007), male gender (p < 0.001),
unfavorable prognostic features (p < 0.001),
treatment with EBVP (epirubicin, bleomycin,
vinblastine, prednisone) plus IFRT (p = 0.002),
and mantle field irradiation alone (p = 0.003).
Therefore, excess mortality was linked to treatment modalities that were associated with poor
failure-free survival resulting in a higher need for
salvage treatment. Late treatment effects are covered in more detail in Chaps. 26–29.
11.5
Combined Modality
Treatment
With the observation of high relapse rates and
fatal long-term effects, most study groups abandoned STNI and EFRT from the treatment of
early-stage HL. Studies were developed in an
Treatment of Early Favorable Hodgkin Lymphoma
Fig. 11.2 The actuarial
risks of death from
major disease categories
in 1261 Dutch HL
patients. Data from
Dutch database on
Hodgkin lymphoma
(Reprinted from Aleman
et al. [25] with
permission)
225
50
Cumulative risk of death (%)
11
Cause of death
40
Hodgkin’s
disease
30
Other than HD
20
Second cancer
10
Cardiovascular
disease
0
0
5
10
15
20
30
25
35
Follow-up (years)
Numbers at risk
1,032
918
attempt to reduce long-term toxicity without
increasing disease-specific mortality. Most randomized studies evaluated CMT in an attempt to
define the optimal chemotherapy, number of
cycles needed, as well as radiation field size and
dose when combined with chemotherapy.
Commonly used regimen and drug combinations
are listed in Table 11.3.
796
High relapse rates after treatment with radiotherapy alone prompted several groups to study CMT
as induction therapy. An earlier meta-analysis of
individual patient data showed that CMT reduced
the relapse risk compared with radiotherapy
alone, but did not improve overall survival [23].
Most of the trials included in this analysis were
conducted between 1967 and 1988 using MOPP
or MOPP-like regimens, which produced unacceptable hematologic toxicity, frequently induced
secondary malignancies, and rendered most
recipients infertile. These studies are therefore
only of historical interest and will not be discussed further. Later, based mainly on results of
studies in advanced HL, the ABVD regimen
became the standard of care in early favorable
272
100
Table 11.3 Chemotherapy regimens used in early-stage
favorable HL
Regimen
ABVD
EBVP
MOPP
MOPP–
ABV
11.5.1 Radiotherapy Alone Versus
CMT
513
Stanford
V
VBM
Drug combinations
Doxorubicin, vinblastine, bleomycin,
dacarbazine
Epirubicin, bleomycin, vinblastine,
prednisone
Mechlorethamine, vincristine,
procarbazine, prednisone
Mechlorethamine, vincristine,
procarbazine, prednisone, doxorubicin,
bleomycin, vinblastine
Vinblastine, doxorubicin, vincristine,
bleomycin, mechlorethamine, etoposide,
prednisone
Vinblastine, methotrexate, bleomycin
HL. When compared with MOPP, ABVD had a
better efficacy and produced less toxicity [28]. In
particular, secondary leukemias and infertility
were less frequently observed than after alkylating agent-containing regimens.
Two randomized studies, one in Germany
and one in the United States, showed the benefit
of adjuvant chemotherapy with a short course of
ABVD or ABVD-like chemotherapy in early
favorable patients: the GHSG HD7 trial compared EFRT alone with CMT consisting of two
cycles of ABVD followed by EFRT in early
favorable patients [11]. A significant advantage
W. Plattel and P. Lugtenburg
226
in freedom from treatment failure (FFTF) was
seen after CMT, mainly related to fewer relapses
as compared with EFRT only (3% vs. 22%).
There were no differences in overall survival
between treatment arms. Importantly, CMT was
not associated with significantly more acute or
long-term toxicity. A trial from the United States
confirmed the benefit of a short course of limited chemotherapy added to STNI in clinically
staged IA and IIA patients [29]. The study
showed that three cycles of doxorubicin and
vinblastine (AV) followed by STNI were well
tolerated and gave a superior failure-free survival compared with STNI alone. The conclusion from these two studies is that the number of
relapses can be reduced by the addition of
ABVD or ABVD-like chemotherapy to large
radiation fields. However, long-­
term toxicity
was still of concern due to the use of extensive
radiation fields.
The Group Pierre-et-Marie-Curie showed that
it was possible to replace the classic mantle field
irradiation with a more limited radiotherapy to
initially involved areas only. This novel approach
termed IFRT involved the addition of chemotherapy to control occult disease in uninvolved areas
[31]. IFRT reduced the irradiation of normal tissues, such as the breast, heart, and lungs.
Therefore, several groups performed randomized trials comparing STNI with a combined
modality approach in which patients received
smaller radiation fields and combination chemotherapy. The results of a selection of some of the
largest trials are listed in Table 11.4.
Table 11.4 Early-stage favorable HL: selection of studies comparing STNI alone with combined modality treatment
(CMT)
Trial
SWOG/CALGB
Year
1989–
2000
Study arms
A. STNI (36–40 Gy)
B. 3 AV + STNI
(36–40 Gy)
Stanford–Kaiser
Permanente
1988–
1995
A. STNI (30–44 Gy)
B. 6 VBM + mantle
field RT
EORTC H7F
1988–
1993
A. STNI (36 Gy)
B. 6 EBVP + IFRT
(36 Gy)
EORTC–GELA
H8F
GHSG HD7
1993–
1999
1994–
1998
A. STNI (36 Gy)
B. 3 MOPP–
ABV + IFRT
(36 Gy)
A. EFRT 30 Gy (IFRT
40 Gy)
B. 2 ABVD + EFRT
30 Gy (IFRT
40 Gy)
Number of
patients
Outcome
326
A. 81% FFS
(3 years)
B. 94% FFS
(3 years)
p < 0.001
78
A. 92% PFS
(5 years)
B. 87% PFS
(5 years)
p = 0.73 (NS)
333
A. 78% EFS
(10 years)
B. 88% EFS
(10 years)
p = 0.0113
542
A. 68% EFS
(10 years)
B. 93% EFS
(10 years)
p < 0.001
627
A. 67% FFTF
(7 years)
B. 88% FFTF
(7 years)
p < 0.0001
Overall
survival
Follow-up
too short
A. 98% OS
(5 years)
B. 94% OS
(5 years)
p = 0.05 (NS)
A. 92% OS
(10 years)
B. 92% OS
(10 years)
p = 0.79 (NS)
A. 92% OS
(10 years)
B. 97% OS
(10 years)
p = 0.001
A. 92% OS
(7 years)
B. 94% OS
(7 years)
p = 0.43 (NS)
Reference
Press et al.
[29]
Horning
et al. [30]
Noordijk
et al. [32]
Fermé et al.
[10]
Engert et al.
[11]
SWOG Southwest Oncology Group, CALGB Cancer and Leukemia, EORTC European Organization for Research and
Treatment of Cancer, GELA Groupe d’Etude des Lymphomes de l’Adulte, GHSG German Hodgkin Study Group, STNI
subtotal nodal irradiation, IFRT involved-field radiotherapy, EFRT extended-field radiotherapy, Gy Gray, FFS failure-­
free survival, PFS progression-free survival, EFS event-free survival, FFTF freedom from treatment failure, OS overall
survival, NS not significant
11
Treatment of Early Favorable Hodgkin Lymphoma
227
In the EORTC H7F trial, patients with early
favorable disease were treated with six cycles of
EBVP followed by IFRT or STNI [32]. The
10-year event-free survival rate after EBVP and
IFRT was 10% better than after STNI alone,
whereas overall survival was 92% in both arms.
This trial demonstrated that EFRT could be
replaced by CMT including IFRT. However, in
early unfavorable patients, EBVP was significantly less efficient than MOPP–ABV [32].
In the subsequent H8F trial by the EORTC–
GELA, favorable HL patients were randomized
between STNI or CMT consisting of three cycles
of MOPP–ABV hybrid followed by IFRT [10].
Patients in the CMT arm had a lower relapse rate,
which resulted in a significantly higher event-free
survival rate than for patients in the STNI arm
(93% vs. 68% at 10 years). Importantly, patients
in the combined modality arm also had a significantly higher overall survival (97% vs. 92% at
10 years). The results of this study again demonstrated the superiority of CMT over EFRT alone
and showed that IFRT is a sufficient treatment
after chemotherapy for early favorable
HL. However, due to its carcinogenic potential,
MOPP–ABV was abandoned in favor of ABVD.
difference between 30 and 20 Gy IFRT [33].
Importantly, there was also no significant difference in terms of overall survival, FFTF, and progression-free survival when all four arms were
compared. The results were robust with longer
follow-up (8 years). The treatment arms with four
cycles of ABVD and 30 Gy IFRT showed significantly more acute toxicity in comparison with
two cycles of ABVD and 20 Gy IFRT. Two cycles
of ABVD followed by 20 Gy IFRT are thus
GHSG standard of care for HL patients in early
favorable stages.
11.5.2 O
ptimal Number of Cycles
of Chemotherapy
The use of fewer cycles of ABVD could potentially reduce late side effects of combined modality therapy. Between 1998 and 2003, the GHSG
HD10 trial accrued more than 1300 favorable
prognosis stage I–II HL patients. Patients were
randomized to four arms in a 2 × 2 factorial
design: two cycles of ABVD followed by 30 Gy
IFRT, two cycles of ABVD followed by
20 Gy IFRT, four cycles of ABVD followed
by 30 Gy IFRT, and four cycles of ABVD followed by 20 Gy IFRT. This trial tested a possible
reduction in the number of ABVD cycles as well
as reduction of radiation dose when using
IFRT. With a median follow-up of 90 months,
there were no significant differences in FFTF and
overall survival at 5 years between four or two
cycles of ABVD. In addition, there was also no
11.5.3 Optimal Chemotherapy
Combination
Reduction of chemotherapy-induced toxicity was
pursued in the GHSG HD13 trial. This trial investigated whether drugs can be omitted from the
ABVD regimen and randomized patients with
early favorable HL to two cycles of either ABVD,
AVD, ABV, or AV with all arms followed by
30 Gy IFRT. Compared with ABVD, the 5-year
FFTF was reduced up to 11.7% (ABV) or 16%
(AV) when dacarbazine and dacarbazine and
bleomycin were deleted and reduced up to 3.9%
(AVD) by the deletion of bleomycin. The reduction in FFTF did not translate into poorer OS
[34]. Therefore, it seems that particularly dacarbazine and to a lesser extent bleomycin are relevant therapeutic agents in ABVD. The Stanford
group has reported good results in 87 patients
with stage I or IIA non-bulky HL treated with an
abbreviated Stanford V regimen administered
weekly for 8 weeks followed by 30 Gy modified
IFRT [35]. At a median follow-up of 10 years, the
FFP was 94%.
11.5.4 Optimal Radiation Dose
Apart from the choice of cytostatic agents and the
number of courses, the question of radiation field
size and dose has also been evaluated (for a selection of randomized trials, see Table 11.5). A
decline in late complications is expected with
lower radiation doses as their incidence is directly
W. Plattel and P. Lugtenburg
228
Table 11.5 Early-stage favorable HL: selection of studies of RT field size and dose in CMT
Trial
Milan
Year
1990–
1997
EORTC–
GELA H9F
1998–
2004
Study arms
A. 4 ABVD + STNI
36–40 Gy
B. 4 ABVD + IFRT
36–40 Gy
A. 6 EBVP + IFRT
36 Gy
B. 6 EBVP +
IFRT20 Gy
C. 6 EBVP (no RT)
Number
of
patients
133
783
Median follow-up
33 months
GHSGHD10
1998–
2003
A. 2 ABVD + IFRT
30 Gy
B. 2 ABVD + IFRT
20 Gy
C. 4 ABVD + IFRT
30 Gy
D. 4 ABVD + IFRT
20 Gy
Median follow-up
91 months
1.370
Outcome
A. FFP 93%
(12 years)
B. FFP 94%
(12 years)
A. EFS 87%
(4 years)
B. EFS 84%
(4 years)
C. EFS 70%
(4 years)
No RT arm closed
because of excess
failure rate
(p < 0.001)
No differences in
FFTF between
patients given two
or four cycles of
ABVD or 20 or
30 Gy IFRT (FFTF
91–93%)
Overall survival
A. OS 96%
(12 years)
B. OS 94%
(12 years)
A. OS 98%
(4 years)
B. OS 98%
(4 years)
C. OS 98%
(4 years)
Reference
Bonadonna
et al. [36]
No survival
differences
between patients
given two or four
cycles of ABVD or
20 or 30 Gy IFRT
(OS 96–97%)
Engert et al.
[33]
Noordijk
et al. [37]
EORTC European Organization for Research and Treatment of Cancer, GELA Groupe d’Etude des Lymphomes de
l’Adulte, GHSG German Hodgkin Study Group, STNI subtotal nodal irradiation, IFRT involved-field radiotherapy, RT
radiotherapy, Gy Gray, FFP freedom from progression, OS overall survival, EFS event-free survival, FFTF freedom
from treatment failure
correlated with the dose of radiation
administered.
Two randomized trials have investigated
radiation doses in early favorable HL patients
treated with CMT. In the EORTC–GELA H9F
trial, 783 patients with stage I–II disease and
favorable characteristics received six cycles of
EBVP. Patients in complete remission after
chemotherapy were randomized to receive
standard dose IFRT (36 Gy), low-dose IFRT
(20 Gy), or no RT at all. The experimental arm
without RT was closed early due to an excess
failure rate compared with the two RT arms,
but there were no differences in outcome
reported between the two radiation dose
levels.
As discussed in Sect. 11.5.2, the GHSG HD10
trial compared doses of 30 and 20 Gy IFRT after
two or four cycles of ABVD. No significant dif-
ferences were observed between patients receiving 30 Gy IFRT and 20 Gy IFRT in terms of
overall survival (97.7 vs. 97.5%), FFTF (93.4 vs.
92.9%), and progression-free survival (93.7 vs.
93.2%), respectively [33]. Therefore, IFRT with
a dose of 20 Gy seems to be sufficient after two
cycles of ABVD.
11.5.5 Optimal Radiation Field Size
The rationale for reduced radiation therapy field
size is to decrease potential late complications
such as cardiovascular and secondary cancers as
the amount of irradiated normal tissue is reduced.
Several randomized trials in early unfavorable
HL have shown that after effective chemotherapy,
IFRT is as effective as EFRT in terms of overall
survival and FFTF [10, 38]. However, data from
11
Treatment of Early Favorable Hodgkin Lymphoma
229
randomized trials in patients with early favorable
HL are scarce.
Bonadonna et al. reported the long-term follow-­up of 133 patients with early HL randomly
assigned to IFRT or STNI after four cycles of
ABVD and found no significant differences in
overall survival (94 vs. 96%) or freedom from
progression (94 vs. 93%) at 12 years [36] (see
Table 11.5). The limited size of the patient sample, however, had no adequate statistical power to
test for non-inferiority of IFRT vs. STNI.
The EORTC–GELA group introduced the
concept of involved-node radiotherapy (INRT)
to further decrease the radiotherapy fields [39,
40]. INRT only includes the initially involved
lymph nodes with a small isotropic margin.
Identifying and contouring involved lymph
nodes is of outmost importance. Therefore, it is
recommended that all patients have cervical and
thoracic CT scans pre- and post-chemotherapy,
preferably in the radiotherapy position, and must
be examined by the radiation oncologist before
the start of the chemotherapy [39, 41]. Better
sparing of normal tissues such as the salivary
glands, heart, coronary arteries, and breast in
female patients is expected with the use of INRT
compared to IFRT (Fig. 11.3). The new INRT
concept was applied in the EORTC–GELA–FIL
H10 randomized trial for patients with earlystage HL. As is shown later in this chapter, INRT
was associated with higher PFS rates compared
to no radiotherapy.
a
c
b
d
Fig. 11.3 Comparison between radiation field sizes and the volume of heart irradiation using either IFRT (a, b) or
INRT (c, d) for a mediastinal tumor mass (PTV in red color) (Reprinted from Girinsky et al. [39] with permission)
W. Plattel and P. Lugtenburg
230
Canadian researchers reported promising
results with INRT in a retrospective study,
although a greater radiation margin was applied
as in the EORTC–GELA–FIL H10 trial [42]. In
British Columbia, the extent of the radiation therapy field size underwent serial changes during
the last decades, from EFRT to IFRT and eventually since 2001 to INRT with margins from 1.5 to
5 cm. There were no statistically significant differences among the three groups for PFS and
overall survival. There were also no marginal
recurrences in the INRT patient group [42].
Although the exact definition of INRT needs further standardization, the concept of INRT seems
feasible.
11.6
Chemotherapy Alone
The potentially life-threatening late side effects
of radiotherapy for HL patients have raised the
question whether those in early-stage disease can
be treated with chemotherapy alone. This question is particularly relevant for patients in whom
the risk of RT-induced toxicity is deemed less
acceptable. Chemotherapy-only protocols have
been successfully used in children and adolescents (see Chap. 14 on pediatric HL). However,
few data exist on their role in adults. Table 11.6
shows a selection of randomized trials performed
in adult patients with early favorable HL dealing
with the issue of chemotherapy alone. These trials encountered a number of problems with
design, patient accrual, as well as variations in
the type of chemotherapy and field size of radiation therapy utilized.
The use of chemotherapy alone is not a new
concept. In the early 1990s, two trials comparing
MOPP with radiotherapy as first-line therapy in
early-stage HL were published [43, 44]. Long
relapse-free survival varied from 70% to 80%,
with varying outcomes of salvage chemotherapy.
The National Cancer Institute of Canada
(NCI-C) and the Eastern Cooperative Oncology
Group (ECOG) conducted a randomized phase
III trial addressing the role of ABVD alone for
early favorable and unfavorable HL. The experimental arm consisted of four cycles of ABVD
alone if a complete remission was achieved after
two cycles. Otherwise, patients received six
cycles. The standard arm was STNI with 36 Gy.
Among the favorable-risk patients, there was no
difference between the two arms for event-free
survival, freedom from disease progression, and
overall survival after a median follow-up of
11.3 years [45]. However, even longer follow-up
is still needed to determine late toxicities.
Only two randomized trials comparing CMT
with chemotherapy alone in early favorable
patients have been published. As discussed in
Sect. 11.5.4, one was the EORTC–GELA H9F
trial in which IFRT in 36 Gy was compared with
20 Gy or no radiotherapy in CR patients after six
cycles of EBVP. The chemotherapy-only arm
was prematurely closed due to an excessive number of relapses [37].
The Memorial Sloan Kettering Cancer Center
randomized early non-bulky HL patients between
six cycles of ABVD alone and six cycles of
ABVD plus 36 Gy radiotherapy. Due to the poor
accrual rate, the trial was closed before completion, and only 152 patients were randomized. No
significant differences were observed between
CMT and chemotherapy alone, but the sample
size was insufficient [46].
11.7
reatment Adaptation Based
T
on PET Scan Response
Functional imaging with PET scanning has
become the standard tool for staging and response
assessment in HL (see Chap. 7). Interim PET
scanning enables evaluation of early metabolic
changes rather than the morphologic changes
occurring later during and after treatment. Several
studies using PET after two or three cycles of
ABVD have shown that early metabolic changes
are predictive of the final treatment response and
PFS [47–51]. Based on these studies which were
mainly performed among advanced stage
patients, several cooperative groups incorporated
interim PET imaging in their early-stage trials to
reduce treatment exposure in responding patients
to prevent overtreatment and/or intensify treatment in case of nonresponsiveness [52–54]. A
11
Treatment of Early Favorable Hodgkin Lymphoma
231
Table 11.6 Early-stage favorable HL: selection of randomized studies of chemotherapy alone in adult patients
Trial
NCI-US
Year
1978–
1989
Study arms
A. 6–8 MOPP
B. Radiotherapy
Rome–
Florence
NCI-C/ECOG
HD6
1979–
1982
1994–
2002
A. Mantle
field + Para-aortic
RT (36–44 Gy)
B. 6 MOPP
A. 4–6 ABVD
Number
of patients Outcome
84
A. DFS 82%
(10 years)
B. DFS 67%
(10 years)
p = NS
89
A. RFS 70% (8 years)
B. RFS 71% (8 years)
123
B. STNI
EORTC–
GELA H9F
1998–
2004
A. 6 EBVP + IFRT
36 Gy
B. 6 EBVP + IFRT
20 Gy
C. 6 EBVP (no RT)
B. EFS 88% (5 years)
783
1990–
2000
A. 6 × ABVD
p = 0.6 (NS)
A. EFS 87% (4 years)
B. EFS 84% (4 years)
Median follow-up
33 months
Memorial
Sloan
Kettering
Cancer center
p = NS
A. EFS 87% (5 years)
152
B. 6 × ABVD + RT
Overall survival
A. OS 90%
(10 years)
B. OS 85%
(10 years)
p = NS
A. OS 93%
(8 years)
B. OS 56%
(8 years)
p < 0.001
A. OS 97%
(5 years)
B. OS 100%
(5 years)
p = 0.3 (NS)
A. OS 98%
(4 years)
B. OS 98%
(4 years)
C. OS 98%
(4 years)
C. EFS 70% (4 years)
No RT arm closed
because of excess
failure rate (p < 0.001)
A. FFP 81% (5 years) A. OS 90%
(5 years)
B. FFP 86% (5 years) B. OS 97%
(5 years)
p = 0.61 (NS)
p = 0.08 (NS)
Reference
Longo
et al. [43]
Biti et al.
[44]
Meyer
et al. [45]
Noordijk
et al. [37]
Strauss
et al. [46]
NCI-US National Cancer Institute United States, EORTC European Organization for Research and Treatment of Cancer,
NCI-C National Cancer Institute of Canada, ECOG Eastern Cooperative Oncology Group, GELA Groupe d’Etude des
Lymphomes de l’Adulte, STNI subtotal nodal irradiation, IFRT involved-field radiotherapy, RT radiotherapy, Gy Gray,
NS not significant, FFP freedom from progression, OS overall survival, DFS disease-free survival, RFS relapse-free
survival, EFS event-free survival
summary of the results of these large randomized
trials are displayed in Table 11.7 and Fig. 11.4.
In the NCI rapid trial, patients with stage IA or
IIA non-bulky HL were treated with three cycles
of ABVD after which PET scanning was performed. The PET scan was negative in 426 out of
602 patients (75%). These 426 patients were randomized between no further treatment and
IFRT. In the intention-to-treat analysis, the PFS
after 3 years in the no further treatment arm was
90.8% versus 94.6% in the IFRT arm. Because of
the large numbers of cross-overs in this trial, the
per protocol analysis is also of interest. In this
analysis, PFS was 90.8% versus 97.1% in the arm
including IFRT [52].
The EORTC/GELA/FIL H10 trial with a total
of 1950 randomized patients also investigated
chemotherapy-only strategies in case of interim
PET negativity. In this trial, patients with early
favorable disease were treated in the standard
arm with three cycles of ABVD and 30 Gy
INRT. An interim PET scan was performed after
two cycles, but no treatment change was performed on the basis of this scan. In the experimental arm, there was both a de-escalation
non-inferiority question in case of a negative
interim PET scan and an escalating superiority
question in case of a positive interim PET scan.
In patients with negative PET findings, INRT was
substituted by a single extra cycle of ABVD in
W. Plattel and P. Lugtenburg
232
Table 11.7 Results of PET response-adapted early favorable Hodgkin lymphoma trials focusing at de-escalation by
leaving out radiotherapy: EORTC H10-F, UK RAPID and GHSG HD16trials
Trial
UK
RAPID
trial
Year
2003–
2010
EORTC
H10-F
2006–
2011
GHSG
HD16
2009–
2015
Study arms
3 ABVD if PET
negative followed by:
A. 30 Gy IF-RT
B. No further
treatment
2 ABVD if PET
negative followed by:
A. 3×ABVD + IN-RT
B. 4×ABVD (no
radiotherapy).
2 ABVD if PET
negative followed by:
A. 20 Gy IF-RT
B. No further
treatment
Number of
patients
426
Overall
survival
At 3 years
A. 97.1%
B. 99%.
Outcome
PFS at 3 years
(per protocol
analysis)
A. 97.1%
B. 90.8%.
PFS at 5 years:
A. 99%
B. 87.1%
1950
1150
PFS at 5 years:
A. 93.4%
B. 86.1%.
Reference
Radford et al.
[52]
At 5 years:
A. 100
B. 99.6%.
Andre et al.
[53]
At 5 years:
A. 98.1%
B. 98.4%
Fuchs et al.
[54]
PFS progression-free survival, OS overall survival, INRT involved-node radiotherapy, IFRT involved-field radiotherapy,
n.a. not available
a
b
NCI UK RAPID
90
Radiotherapy
80
No further treatment
90
70
60
50
40
30
20
Rate ratio, 1.57 (95% CI, 0.84-2.97)
P=0.16
10
0
0
12
24
36
48
60 72
84
96 108 120
c
80
Chemotherapy only
70
60
50
40
30
20
10
0
Months since Randomization
PFS Rate (%)
Radiotherapy
100
Progression free survival (%)
Progression free survival (%)
100
EORTC/FIL/LYSA H10
HR, 15.8 (95% CI, 3.79 to 66.07)
1
2
3
4
5
6
7
8
Time (years)
GHSG HD16
1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
5-year estimate (95% CI)
2x ABVD + 20 Gy IFRT 93.4% (90.4% to 96.5%)
2x ABVD
86.1% (81.4% to 90.9%)
–7.3% (–13.0% to –1.6%)
Difference
0
Hazard ratio (95% CI)
Log-rank test
1.78 (1.02 to 3.12)
P = .040
Median follow-up
47 months
12
36
24
Time (months)
48
60
Fig. 11.4 Results of the recently published trials of PET
adapted trials in early favorable Hodgkin lymphoma.
Shown are progression-free survival curves for the UK
RAPID trial (a), EORTC/LYSA/FIL H10 favorable (b)
and GHSG HD16 (c)
11
Treatment of Early Favorable Hodgkin Lymphoma
b
100
90
escBEACOPP + INRT
80
70
ABVD + INRT
60
50
40
30
20
10
HR, 0.42 (95% CI, 0.23 to 0.74); P = .002
1
2
3
4
5
6
7
escBEACOPP + INRT
100
90
Overall survival (%)
Progression free survival (%)
a
233
8
Time (years)
ABVD + INRT
80
70
60
50
40
30
20
10
HR, 0.45 (95% CI, 0.19 to 1.07); P = .062
1
2
3
4
5
6
7
8
Time (years)
Fig. 11.5 Results of the EORTC/LYSA/FIL H10 trial in
case of a positive FDG-PET scan after 2 ABVD among
patients with both favorable and unfavorable early-stage
HL. Shown are progression-free survival (a) and overall
survival (b)
favorable subgroup patients and two extra cycles
in patients with unfavorable disease. The de-­
escalation arm with the substitution of radiotherapy by extra chemotherapy was closed
prematurely due to futility based on 33 events. In
line with the results of the RAPID trial, final
results at 5-year follow-up showed that early
favorable patients with a negative PET after two
cycles of ABVD have an excellent outcome when
treated with CMT (5-year PFS 99%). Substituting
radiotherapy by a single extra course of ABVD
resulted in a decrease of about 12% PFS [53].
Similar were reported from GHSG HD16 trial
which was recently published. In this large trial
involving 1150 patients with early favorable HL,
patients with a negative PET scan after two cycles
of ABVD were randomized between standard
30 Gy IFRT and no further treatment. Again
omission of radiotherapy resulted in a decrease of
tumor control with PFS of 93.4% at 5 years in the
CMT arm versus 86.1% in the chemotherapyonly arm [54]. There were no differences in overall survival at 5 years.
Taken together, these three trials demonstrated that omission of radiotherapy among
patients with early-stage HL and a negative PET
scan after two or three cycles of ABVD resulted
in a clinically relevant decrease in PFS. Although
it must be mentioned that chemotherapy-only
strategies based on interim PET scanning also
resulted in excellent treatment outcomes and
can be seen as a treatment option for patients in
whom radiotherapy is expected to result in
excessive short- and/or long-term toxicity.
Overall survival in all trials was not different
between treatment arms meaning that almost all
patients not receiving radiotherapy could successfully be salvaged. Long-term effects on
overall survival of the omission of radiotherapy
and the impact of salvage treatments need to be
awaited.
The EORTC H10 study was the only study
that also investigated escalation of treatment
based on a positive interim PET scan. Patients
with both early favorable and early unfavorable
HL and a positive interim PET scan after two
cycles of ABVD subsequently received two
cycles of escBEACOPP. In this joint group, escalation to escBEACOPP + INRT resulted in
improved 5-year PFS of 90.6% compared to
77.4% for standard three cycles (early favorable)
or four cycles (early unfavorable) of
ABVD + INRT (Fig. 11.5) [53].
In conclusion, interim PET-guided treatment
results in improved tumor control in patients with
positive findings by escalating chemotherapy on
one hand, and it gives the possibility to relatively
safely omit radiotherapy where needed in patients
with a complete response on ABVD chemotherapy on the other hand.
W. Plattel and P. Lugtenburg
234
11.8
Recommendations
and Future Directions
In most parts of the world, CMT strategies including 2–3 cycles of ABVD followed by 20–30 Gy
IFRT will remain standard treatment for patients
with early favorable HL. Incorporation of an
interim PET-guided approach with escalation to
escBEACOPP in case of positive findings might
further improve the already excellent treatment
results. In patients at increased risk of RT-related
toxicity because of age, sex, and disease localization, PET-guided treatment might aid the selection of patients in whom radiotherapy can be
relatively safely omitted. Although, in such cases,
omission of radiotherapy will result in a decrease
of local tumor control of about 7–12%, outcomes
with chemotherapy only are still excellent.
Therefore, balancing the risk of RT-related toxicity to the possible decrease in local tumor control
is the main challenge when planning treatment
upfront. When balancing this risk, it is important
to realize that data on toxicity of radiotherapy are
mainly based on past radiotherapy techniques,
fields, and doses, and all have been massively
improved last decades. It is therefore of outmost
importance to collect long-term follow-up outcomes of current treatment modalities.
With this approach, PFS rates exceeding 90%
and OS rates exceeding 95% can be achieved. At
present, the goal in early favorable HL is to maintain the excellent efficacy while further reducing
acute and late toxicity. Incorporation or replacing
current chemotherapy regimens by successful
new drugs in HL like brentuximab vedotin or
checkpoint inhibitors might be a further improvement and a possible route to chemotherapy free
treatment in HL. However, it might be difficult to
improve on current excellent treatment results
with only short courses of limited chemotherapy
or radiotherapy. Other opportunities to improve
outcome of early favorable HL are the
­introduction of proton beam radiotherapy, better
selection of patients for certain treatments based
on biomarkers, or better methods for detection of
minimal residual disease (MRD) in HL. These
efforts are currently being made and are further
discussed in Chap. 13.
References
1. Carbone PP, Kaplan HS, Musshoff K, Smithers
DW, Tubiana M (1971) Report of the committee on
Hodgkin’s disease staging classification. Cancer Res
31:1860–1861
2. Lister TA, Crowther D, Sutcliffe SB, Glatstein E,
Canellos GP, Young RC et al (1989) Report of a committee convened to discuss the evaluation and staging
of patients with Hodgkin’s disease: Cotswolds meeting. J Clin Oncol 7:1630–1636
3. Juweid ME, Stroobants S, Hoekstra OS, Mottaghy
FM, Dietlein M, Guermazi A et al (2007) Use of positron emission tomography for response assessment of
lymphoma: consensus of the imaging Subcommittee
of International Harmonization Project in lymphoma.
J Clin Oncol 25:571–578
4. Cheson BD, Pfistner B, Juweid ME, Gascoyne RD,
Specht L, Horning SJ et al (2007) Revised response
criteria for malignant lymphoma. J Clin Oncol
25:579–586
5. El-Galaly TC, d’Amore F, Mylam KJ, de Nully
Brown P, Bøgsted M, Bukh A et al (2012) Routine
bone marrow biopsy has little or no therapeutic consequence for positron emission tomography/computed
tomography-­
staged treatment naïve patients with
Hodgkin lymphoma. J Clin Oncol 30:4508–4514
6. Specht L, Nissen NI (1988) Hodgkin’s disease stages
I and II with infradiaphragmatic presentation: a rare
and prognostically unfavourable combination. Eur J
Haematol 40:396–402
7. Sasse S, Goergen H, Plütschow A, Böll B, Eichenauer
DA, Fuchs M et al (2018) Outcome of patients with
early-stage infradiaphragmatic Hodgkin lymphoma:
a comprehensive analysis from the German Hodgkin
study group. J Clin Oncol 36:2603–2611
8. Specht L (1996) Prognostic factors in Hodgkin’s disease. Semin Radiat Oncol 6:146–161
9. Tubiana M, Henry-Amar M, van der Werf-Messing
B, Henry J, Abbatucci J, Burgers M et al (1985) A
multivariate analysis of prognostic factors in early
stage Hodgkin’s disease. Int J Radiat Oncol Biol Phys
11:23–30
10. Ferme C, Eghbali H, Meerwaldt JH, Rieux C, Bosq
J, Berger F et al (2007) Chemotherapy plus involved-­
field radiation in early stage Hodgkin’s disease. N
Engl J Med 357:1916–1927
11. Engert A, Franklin J, Eich HT, Brillant C, Sehlen
S, Cartoni C et al (2007) Two cycles of doxorubicin, bleomycin, vinblastine, and dacarbazine plus
extended-field radiotherapy is superior to radiotherapy alone in early favorable Hodgkin’s lymphoma:
final results of the GHSG HD7 trial. J Clin Oncol
25:3495–3502
12. Cottereau AS, Versari A, Loft A, Casasnovas O, Bellei
M, Ricci R et al (2018) Prognostic value of baseline
metabolic tumor volume in early-stage Hodgkin lymphoma in the standard arm of the H10 trial. Blood
131:1456–1463
11
Treatment of Early Favorable Hodgkin Lymphoma
13. Hoppe RT, Coleman CN, Cox RS, Rosenberg SA,
Kaplan HS (1982) The management of stage I–II
Hodgkin’s disease with irradiation alone or combined
modality therapy: the Stanford experience. Blood
59:455–465
14. Gospodarowicz MK, Sutcliffe SB, Clark RM, Dembo
AJ, Fitzpatrick PJ, Munro AJ et al (1992) Analysis of
supradiaphragmatic clinical stage I and II Hodgkin’s
disease treated with radiation alone. Int J Radiat
Oncol Biol Phys 22:859–865
15. Raemaekers J, Kluin-Nelemans H, Teodorovic I,
Meerwaldt C, Noordijk E, Thomas J et al (2002)
The achievements of the EORTC lymphoma group.
European Organisation for Research and Treatment of
Cancer. Eur J Cancer 38(Suppl 4):S107–S113
16. Tubiana M, Henry-Amar M, Carde P, Burgers JM,
Hayat M, Van der Schueren E et al (1989) Toward
comprehensive management tailored to prognostic factors of patients with clinical stages I and II in
Hodgkin’s disease. The EORTC lymphoma group
controlled clinical trials: 1964–1987. Blood 73:47–56
17. Tubiana M, Henry-Amar M, Hayat M, Breur K, van
der Werf-Messing B, Burgers M (1979) Long-term
results of the E.O.R.T.C. randomized study of irradiation and vinblastine in clinical stages I and II of
Hodgkin’s disease. Eur J Cancer 15:645–657
18. Tubiana M, Hayat M, Henry-Amar M, Breur K, van
der Werf MB, Burgers M (1981) Five-year results
of the E.O.R.T.C. randomized study of splenectomy
and spleen irradiation in clinical stages I and II of
Hodgkin’s disease. Eur J Cancer 17:355–363
19. Carde P, Burgers JM, Henry-Amar M, Hayat M,
Sizoo W, Van der Schueren E et al (1988) Clinical
stages I and II Hodgkin’s disease: a specifically tailored therapy according to prognostic factors. J Clin
Oncol 6:239–252
20. Carde P, Hagenbeek A, Hayat M, Monconduit M,
Thomas J, Burgers MJ et al (1993) Clinical staging versus laparotomy and combined modality
with MOPP versus ABVD in early stage Hodgkin’s
disease: the H6 twin randomized trials from the
European Organization for Research and Treatment
of Cancer lymphoma cooperative group. J Clin Oncol
11:2258–2272
21. Noordijk EM, Carde P, Dupouy N, Hagenbeek A,
Krol AD, Kluin-Nelemans JC et al (2006) Combined-­
modality therapy for clinical stage I or II Hodgkin’s
lymphoma: long-term results of the European
Organisation for Research and Treatment of Cancer
H7 randomized controlled trials. J Clin Oncol
24:3128–3135
22. Duhmke E, Franklin J, Pfreundschuh M, Sehlen S,
Willich N, Ruhl U et al (2001) Low-dose radiation
is sufficient for the noninvolved extended-field treatment in favorable early stage Hodgkin’s disease: long-­
term results of a randomized trial of radiotherapy
alone. J Clin Oncol 19:2905–2914
23. Specht L, Gray RG, Clarke MJ, Peto R (1998)
International Hodgkin’s disease collaborative group.
Influence of more extensive radiotherapy and adju-
235
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
vant chemotherapy on long-term outcome of early
stage Hodgkin’s disease: a meta-analysis of 23 randomized trials involving 3,888 patients. J Clin Oncol
16:830–843
Horwich A, Specht L, Ashley S (1997) Survival analysis of patients with clinical stages I or II Hodgkin’s
disease who have relapsed after initial treatment with
radiotherapy alone. Eur J Cancer 33:848–853
Aleman BM, van den Belt-Dusebout AW, Klokman
WJ, Van’t Veer MB, Bartelink H, van Leeuwen
FE (2003) Long-term cause-specific mortality of
patients treated for Hodgkin’s disease. J Clin Oncol
21:3431–3439
Ng AK, Bernardo MP, Weller E, Backstrand KH,
Silver B, Marcus KC et al (2002) Long-term survival
and competing causes of death in patients with early
stage Hodgkin’s disease treated at age 50 or younger.
J Clin Oncol 20:2101–2108
Favier O, Heutte N, Stamatoullas-Bastard A, Carde
P, Van’t Veer MB, Aleman BM et al (2009) Survival
after HL: causes of death and excess mortality in
patients treated in 8 consecutive trials. Cancer 115:
1680–1691
Canellos GP, Anderson JR, Propert KJ, Nissen N,
Cooper MR, Henderson ES et al (1992) Chemotherapy
of advanced Hodgkin’s disease with MOPP, ABVD,
or MOPP alternating with ABVD. N Engl J Med
327:1478–1484
Press OW, LeBlanc M, Lichter AS, Grogan TM,
Unger JM, Wasserman TH et al (2001) Phase III randomized intergroup trial of subtotal lymphoid irradiation versus doxorubicin, vinblastine, and subtotal
lymphoid irradiation for stage IA to IIA Hodgkin’s
disease. J Clin Oncol 19:4238–4244
Horning SJ, Hoppe RT, Mason J, Brown BW, Hancock
SL, Baer D et al (1997) Stanford-Kaiser Permanente
G1 study for clinical stage I to IIA Hodgkin’s disease: subtotal lymphoid irradiation versus vinblastine, methotrexate, and bleomycin chemotherapy and
regional irradiation. J Clin Oncol 15:1736–1744
Zittoun R, Audebert A, Hoerni B, Bernadou A, Krulik
M, Rojouan J et al (1985) Extended versus involved
fields irradiation combined with MOPP chemotherapy in early clinical stages of Hodgkin’s disease. J
Clin Oncol 3:207–214
Noordijk EM, Carde P, Dupouy N, Hagenbeek A,
Krol AD, Kluin-Nelemans JC et al (2006) Combined-­
modality therapy for clinical stage I or II Hodgkin’s
lymphoma: long-term results of the European
Organisation for Research and Treatment of Cancer
H7 randomized controlled trials. J Clin Oncol
24:3128–3135
Engert A, Plütschow A, Eich HT, Lohri A, Dörken B,
Borchmann P et al (2010) Reduced treatment intensity in patients with early-stage Hodgkin’s lymphoma.
N Engl J Med 363:640–652
Behringer K, Görgen H, Hitz F, Zijlstra JM, Greil R,
Markova J et al (2015) Omission of dacarbazine or
bleomycin, or both, from the ABVD regimen in treatment of early-stage favourable Hodgkin’s lymphoma
W. Plattel and P. Lugtenburg
236
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
(GHSG HD13): an open-label, randomised, non-inferiority trial. Lancet 11;385:1418–1427
Advani RH, Hoppe RT, Baer D, Mason J, Warnke R,
Allen J et al (2013) Efficacy of abbreviated Stanford
V chemotherapy and involved-field radiotherapy in
early-stage Hodgkin lymphoma: mature results of the
G4 trial. Ann Oncol 24:1044–1048
Bonadonna G, Bonfante V, Viviani S, Di Russo A,
Villani F, Valagussa P (2004) ABVD plus subtotal
nodal versus involved-field radiotherapy in early stage
Hodgkin’s disease: long-term results. J Clin Oncol
22:2835–2841
Thomas J, Fermé C, Noordijk EM, Morschhauser
F, Girinsky T, Gaillard I et al (2018) Comparison of
36 Gy, 20 Gy, or no radiation therapy after 6 cycles
of EBVP chemotherapy and complete remission in
early-­
stage Hodgkin lymphoma without risk factors: results of the EORT-GELA H9-F intergroup
randomized trial. Int J Radiat Oncol Biol Phys
100:1133–1145
Engert A, Schiller P, Josting A, Herrmann R, Koch
P, Sieber M et al (2003) Involved-field radiotherapy
is equally effective and less toxic compared with
extended-field radiotherapy after four cycles of chemotherapy in patients with early stage unfavorable
Hodgkin’s lymphoma: results of the HD8 trial of the
German Hodgkin’s lymphoma study group. J Clin
Oncol 21:3601–3608
Girinsky T, van der Maazen R, Specht L, Aleman B,
Poortmans P, Lievens Y et al (2006) Involved-node
radiotherapy (INRT) in patients with early HL: concepts and guidelines. Radiother Oncol 79:270–277
Shahidi M, Kamangari N, Ashley S, Cunningham D,
Horwich A (2006) Site of relapse after chemotherapy
alone for stage I and II Hodgkin’s disease. Radiother
Oncol 78:1–5
Girinsky T, Specht L, Ghalibafian M, Edeline V,
Bonniaud G, Van Der Maazen R et al (2008) The
conundrum of HL nodes: to be or not to be included
in the involved node radiation fields. The EORTC-­
GELA lymphoma group guidelines. Radiother Oncol
88:202–210
Campbell BA, Voss N, Pickles T, Morris J, Gascoyne
RD, Savage KJ et al (2008) Involved-nodal radiation
therapy as a component of combination therapy for
limited-stage Hodgkin’s lymphoma: a question of
field size. J Clin Oncol 26:5170–5174
Longo DL, Glatstein E, Duffey PL, Young RC,
Hubbard SM, Urba WJ et al (1991) Radiation therapy versus combination chemotherapy in the treatment of early stage Hodgkin’s disease: seven-year
results of a prospective randomized trial. J Clin Oncol
9:906–917
Biti GP, Cimino G, Cartoni C, Magrini SM, Anselmo
AP, Enrici RM et al (1992) Extended-field radiotherapy is superior to MOPP chemotherapy for the treat-
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
ment of pathologic stage I–IIA Hodgkin’s disease:
eight-year update of an Italian prospective randomized study. J Clin Oncol 10:378–382
Meyer RM, Gospodarowicz MK, Connors JM,
Pearcey RG, Wells WA, Winter JN et al (2012) ABVD
alone versus radiation-based therapy in limited-stage
Hodgkin’s lymphoma. N Engl J Med 366:399–408
Straus DJ, Portlock CS, Qin J, Myers J, Zelenetz AD,
Moskowitz C et al (2004) Results of a prospective
randomized clinical trial of doxorubicin, bleomycin,
vinblastine, and dacarbazine (ABVD) followed by
radiation therapy (RT) versus ABVD alone for stages
I, II, and IIIA nonbulky Hodgkin disease. Blood
104:3483–3489
Gallamini A, Hutchings M, Rigacci L, Specht L, Merli
F, Hansen M et al (2007) Early interim 2-[18F]fluoro-­
2-­
deoxy-D-glucose positron emission tomography
is prognostically superior to international prognostic score in advanced-stage Hodgkin’s lymphoma: a
report from a joint Italian-Danish study. J Clin Oncol
25:3746–3752
Hutchings M, Mikhaeel NG, Fields PA, Nunan T,
Timothy AR (2005) Prognostic value of interim FDG-­
PET after two or three cycles of chemotherapy in
HL. Ann Oncol 16:1160–1168
Hutchings M, Loft A, Hansen M, Pedersen LM,
Buhl T, Jurlander J et al (2006) FDG-PET after two
cycles of chemotherapy predicts treatment failure and
progression-­free survival in HL. Blood 107:52–59
Oki Y, Chuang H, Chasen B, Jessop A, Pan T, Fanale
M et al (2014) The prognostic value of interim positron emission tomography scan in patients with classical Hodgkin lymphoma. Br J Haematol 165:112–116
Gallamini A, Barrington SF, Biggi A, Chauvie S,
Kostakoglu L, Gregianin M (2014) The predictive
role of interim positron emission tomography for
Hodgkin lymphoma treatment outcome is confirmed
using the interpretation criteria of the Deauville five-­
point scale. Haematologica 99:1107–1113
Radford J, Illidge T, Counsell N, Hancock B,
Pettengell R, Johnson P (2015) Results of a trial of
PET-directed therapy for early-stage Hodgkin’s lymphoma. N Engl J Med 372:1598–1607
André MPE, Girinsky T, Federico M, Reman O,
Fortpied C, Gotti M (2017) Early positron emission
tomography response-adapted treatment in stage I
and II Hodgkin lymphoma: final results of the randomized EORTC/LYSA/FIL H10 trial. J Clin Oncol
35:1786–1794
Fuchs M, Goergen H, Kobe C, Kuhnert G, Lohri A,
Greil R et al (2019) Positron emission tomographyguided treatment in early-stage favorable Hodgkin
lymphoma: final results of the international, randomized phase III HD16 trial by the German Hodgkin
Study Group. J Clin Oncol 37:2835–2845
Treatment of Early Unfavorable
Hodgkin Lymphoma
12
Marc P. E. André and Andreas Engert
Contents
12.1
12.1.1
12.1.2
Prognostic Factors Definition New Prognostic Factors 237
237
240
12.2
Long-Term Side Effects 240
12.3
12.3.1
12.3.2
12.3.3
Non-PET-Adapted Treatment Strategies Fields and Dose of Radiotherapy Chemotherapy Chemotherapy Alone 241
241
241
242
12.4
12.4.1
12.4.2
12.4.2.1
12.4.2.2
12.4.2.3
12.4.2.4
PET-Adapted Treatment Strategies Interim PET Clinical Trials Rapid Study H10 Study Other Studies Management of iPET-Positive Patients 242
242
243
243
243
244
244
12.5
ESMO and NCCN Recommendations 244
12.6
12.6.1
12.6.2
New Drugs Brentuximab Vedotin Checkpoint Inhibitors 245
245
245
12.7
Conclusions and Future Strategies 246
References M. P. E. André (*)
Department of Hematology, Université Catholique de
Louvain, CHU UcL Namur, Yvoir, Belgium
e-mail: marc.andre@uclouvain.be
A. Engert
Department of Internal Medicine I, German Hodgkin
Study Group (GHSG), University Hospital of
Cologne, Cologne, Germany
e-mail: a.engert@uni-koeln.de
246
12.1
Prognostic Factors
12.1.1 Definition
The Ann Arbor staging system with the 1989
Cotswolds modifications is still being used
worldwide in patients with Hodgkin lymphoma
(HL) [1]. Modern staging procedures ­recommend
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_12
237
M. P. E. André and A. Engert
238
the routine use of [18F] fluoro-2-deoxy-D-glucose
positron emission tomography-CT scanning
(PET-CT) at diagnosis [2]. With the introduction
of PET-CT scanning at diagnosis, up to 30% of
patients will be upstaged mainly from early to
advanced stages. In addition, the extent of radiation fields in CS I/II disease can be influenced by
identifying additional lesions by PET-CT scanning [2, 3]. Interestingly, when a PET-CT is performed for initial staging, a bone marrow biopsy
is no longer required [4, 5]. In the study by
El-Galaly et al. [5], 18% of patients showed focal
skeletal lesions on PET-CT, but only 6% had positive bone marrow biopsies. None of the patients
would have been allocated to other treatments
based on bone marrow biopsy results. Patients
with early-stage disease rarely have bone marrow
involvement in the absence of a suggestive PET
finding, confirming that, if a PET-CT is performed, a bone marrow aspirate/biopsy is no longer required for the routine evaluation.
Even in stage I/II, the extent of disease varies
substantially requiring a risk-adapted treatment.
In many early-stage patients, mediastinal bulky
disease is present, which has been demonstrated
as prognostically unfavorable. Other poor prognostic clinical factors include higher age,
increased number of involved nodes, and elevated
erythrocyte sedimentation rate (ESR), accompanied by B symptoms. Though slight differences
in definition exist between major cooperative
groups, clinical stage I/II HL patients in Europe
are generally divided into an early favorable and
an early unfavorable (intermediate) subgroup.
Patients in North America presenting with
adverse factors (mainly the presence of bulky disease) are treated like those having stage III–IV
disease; thus, these patients are not included in
clinical trials for stage I/II disease.
The factors used by the European Organization
for Research and Treatment of Cancer (EORTC)
Lymphoma Group/Lymphoma Study Association
(LYSA), the German Hodgkin Study Group
(GHSG), the National Cancer Institute of Canada
(NCIC), and the Eastern Cooperative Oncology
Group (ECOG) are shown in Table 12.1. These
risk factors and the resulting prognostic groups
were originally defined in the context of treatment with extended-field radiotherapy (RT). In a
combined modality setting, the differences in
prognosis between favorable and unfavorable
disease are likely to be smaller. In more recent
series, the treatment was mainly tailored according to the prognostic group. Thus, one would
have anticipated that these prognostic factors
today have less independent prognostic significance. Klimm et al. analyzed the impact of the
three different staging and prognostic subgroup
definitions on the outcome of 1173 early-stage
patients treated homogeneously in the HD10 and
HD11 trials of the GHSG [6]. Figure 12.1 shows
the PFS of these patients related to the GHSG,
EORTC/LYSA, and NCCN prognostic risk factor
score, respectively: all three staging systems identified the unfavorable risk group. Especially tumorspecific risk factors rather than patient-­specific
risk factors such as mediastinal bulk and high
tumor activity were predictive for poor outcome.
Table 12.1 Risk factors according to cooperative treatment groups
Risk factors (RF)
Stages
Favorable
Unfavorable or
intermediate
EORTC/LYSA
A: Mediastinal mass
GHSG
A: Mediastinal mass
B: Age ≥ 50 years
C: ESR ≥ 50 or ESR ≥ 30
with B symptoms
D: ≥ 4 nodal areas
B: Extranodal disease
C: ESR ≥ 50
NCIC/ECOG
A: Histology other
than LP/NS
B: age > 40
C: ESR > 50
D: ≥ 3 nodal areas
D: ≥ 3 sites
I–II without RF
I–II with ≥1 RF and IIB with
C/D without AB
I–II without RF
I–II with ≥1 RF
I–II without RF
I–II with ≥1 RF
EORTC European Organization for Research and Treatment of Cancer, LYSA Lymphoma Study Association, GHSG
German Hodgkin Study Group, NCIC National Cancer Institute of Canada, ECOG Eastern Cooperative Oncology
group, ESR erythrocyte sedimentation rate, LP lymphocyte predominance, NS nodular sclerosis
12
Treatment of Early Unfavorable Hodgkin Lymphoma
GHSG
239
1.0
0.9
0.8
0.7
5 year estimate [95 %-CI]:
95.8 % [94.0 % to 97.6 %]
Favorable
86.4 % [83.7 % to 89.1 %]
Unfavorable
PFS
0.6
0.5
0.3
p = <.001
0.2
0.1
Favorable
0.0
Pts. at risk
favorable
unfavorable
EORTC
–9.4 % [–12.7 % to –6.2 %]
2.61 [1.74 to 3.91]
difference
Hazard ratio
0.4
0
12
24
525
648
504
614
486
568
36
48
60
Time [months]
463
437
390
543
513
475
Unfavorable
72
84
96
302
375
209
247
123
141
72
84
96
282
348
195
232
118
126
72
84
96
291
375
205
245
125
136
1.0
PFS
0.9
0.8
5 year estimate [95 %-CI]:
0.7
Favorable
Unfavorable
94.2 % [92.1 % to 96.4 %]
87.6 % [84.8 % to 90.3%]
difference
Hazard ratio
–6.7 % [–10.2 % to –3.2 % ]
2.10 [1.41 to 3.13]
0.6
0.5
0.4
0.3
p = <.001
0.2
0.1
Favorable
0.0
Pts. at risk
favorable
unfavorable
NCCN
0
12
24
496
589
473
560
455
517
36
48
60
Time [months]
433
406
355
495
470
444
Unfavorable
1.0
0.9
0.8
5 year estimate [95 %-CI]:
PFS
0.7
0.6
0.5
0.4
Favorable
Unfavorable
95.3 % [93.3 % to 97.2 %]
86.7 % [84.0 % to 89.3%]
difference
Hazard ratio
–8.6 % [–11.9 % to –5.3 % ]
2.14 [1.45 to 3.16]
0.3
p = <.001
0.2
0.1
Favorable
0.0
Pts. at risk
favorable
unfavorable
0
12
24
506
645
483
616
467
569
36
48
60
Time [months]
444
416
371
544
516
478
Fig. 12.1 Estimated progression-free survival using
staging definitions of the German Hodgkin Study Group,
the European Organization for Research and Treatment of
Unfavorable
Cancer (EORTC), or National Comprehensive Cancer
Network (NCCN) [6]
M. P. E. André and A. Engert
240
Table 12.2 Multivariate analysis testing total metabolic tumor volume (TMTV), with interim PET response after two
cycles (iPET2) and individual baseline factors, EORTC, GHSG, NCCN staging systems
TMTV tested with:
A. Individual factors
TMTV > 147 cm3
IPET 2
B symptoms
≥ 4 involved sites
M/T ≥ 0.35
B.EORTC
TMTV > 147 cm3
IPET2
Unfavorable EORTC
C.GHSG
TMTV > 147 cm3
IPET2
Unfavorable GHSG
D.NCCN
TMTV > 147 cm3
IPET2
Unfavorable NCCN
PFSa
HR
95% CI
P
PFS (final model)
HR
95% CI
P
3.9
11.0
2.1
2.0
0.8
1.6–9.5
4.8–25.1
0.9–4.8
0.8–5.2
0.3–2.0
0.0032
<0.0001
0.076
0.16
0.65
4.4
10.9
2.0–9.5
4.9–24.4
0.0002
<0.0001
3.5
9.2
3.2
1.6–7.8
4.1–20.6
0.9–11.1
0.0016
<0.0001
0.067
4.4
10.9
2.0–9.5
4.9–24.4
0.0002
<0.0001
4.1
10.6
1.3
1.8–9.3
4.7–23.9
0.4–4.0
0.0006
<0.0001
0.69
4.4
10.9
2.0–9.5
4.9–24.4
0.0002
<0.0001
3.7
10.2
1.8
1.7–8.4
4.5–22.8
0.6–5.7
0.00014
<0.0001
0.30
4.4
10.9
2.0–9.5
4.9–24.4
0.0002
<0.0001
All variables integrated in the Cox model; final model: with significant factors after performing the backward stepwise
Cox model (Adapted from Cottereau, Blood 2018 with permission)
a
In terms of overall survival, the scores reflected
the unfavorable risk profile as well. These data
underline the continued need for identifying a
poor-risk group within the group of stage I/II disease though new risk factors with a higher specificity might be useful.
12.1.2 New Prognostic Factors
Several different prognostic factors adopted so far
are surrogates of the tumor burden. Specht et al.
[7] were the first to demonstrate the strong prognostic impact of tumor burden attempting to estimate tumor volume. This was based on the
categorization of lesion size by physical examination as well as mediastinal and hilar involvement
(chest X-rays) as well as adding grades of all
involved sites. The superiority of tumor burden
over other prognostic factors was further confirmed by Gobbi et al. [8]. More recently, PET-CT
scanning has been used to define the functionally
active tumor volume using total metabolic tumor
volume (TMTV). Cottereau et al. conducted an
analysis on 294 early-stage HL including interim
PET and TMTV in the different prognostic models (EORTC/LYSA, GHSG and NCC) [9]. In this
analysis, only TMTV and interim PET remained
significant (Table 12.2). Although PET-CT is a
tool that allows to refine prognosis and treatment
strategies if there is a certain degree of inaccuracy
in its application. An area of growing interest is
combing PET-CT and biomarkers such as circulating tumor-free DNA. Spina et al. recently demonstrated that this biomarker could identify
residual disease during treatment of disease after
two courses of treatment [10]. Incorporation of
both, PET-CT and cell-free tumor DNA, in our
decision algorithm will possibly profoundly modify the way we use prognostic factors in the future.
12.2
Long-Term Side Effects
The present management of early-stage HL aims
at curing the disease with a specific attention to the
reduction of late effects. The most severe late
effect due to the treatment of HL is secondary cancer. In a recent large study [11] with a median follow-up of 19.1 years, the standardized incidence
12
Treatment of Early Unfavorable Hodgkin Lymphoma
ratio was 4.6 (95% confidence interval (CI), 4.3–
4.9) in the study cohort when compared with the
general population. The risk was still elevated 35
years or more after treatment (SIR, 3.9; 95% CI,
2.8–5.4), and the cumulative incidence of a second
cancer in the study cohort at 40 years was 48.5%
(95% CI, 45.4–51.5). Unfortunately, the cumulative incidence of second solid cancers did not
differ between study periods (1965–1976, 1977–
1988, or 1989–2000) (P = 0.71 for heterogeneity),
suggesting that the efforts made to reduce the burden of treatment did not translate into a reduction
of second cancers. However, the impact of treatment modifications in the last 20 years is not well
known. Also, as the risk is better known, it might
be suggested that well-conducted cancer screening
programs could also reduce the severity of late
malignancies. However, in the study of Baxstrom
et al. [12], many women did not get the appropriate dual screening for breast cancer despite their
increased risk, with only 36.6% of the study sample receiving dual screening. Proper screening
allows detection of secondary breast cancer at earlier stages where treatment can be local, but this
study raised the issue of compliance of this population to cancer screening programs. Finally, cancer screening is not yet possible for thyroid, lung,
and soft tissue cancers.
Cardiovascular and valvular diseases represent another important late effect occurring in
patients receiving mediastinal radiotherapy [13,
14]. The reduction in dose and volume of radiotherapy led to a reduction in these complications.
Nevertheless, radiotherapy may still result in
substantial incidental cardiac exposure if the disease affects the mediastinum.
12.3
Non-PET-Adapted Treatment
Strategies
12.3.1 Fields and Dose
of Radiotherapy
The use of large radiation fields was abandoned after
both, the GHSG HD8 trial [15] and the H8U trial
conducted by the EORTC/LYSA [16]. In HD8,
long-term noninferiority of involved-field radiotherapy (IF-RT) was compared with extended-field
241
RT. With regard to treatment-­associated long-term
toxicity, a non-significant trend towards less secondary neoplasia was observed with IF-RT in the most
recent follow-up analysis (15-year cumulative, 14%
vs. 17%; p = 0.3) [17]. This trend was more pronounced when examining only the incidence of
acute myeloid leukemia or myelodysplastic syndromes (2.4% vs. 0.8%; p = 0.1), but not in nonHodgkin lymphoma (2.6% vs. 2.9%; p = 1.0). In
solid second neoplasia, the trend became more pronounced with longer follow-up but did not meet statistical significance (12% vs. 10.4%; p = 0.7). Due to
the long latency period of second solid neoplasia,
prolonged follow-up is crucial to finally assess the
risk of secondary malignancies with more limited
RT fields.
In the H8U trial, 42 of 766 (5%) patients
relapsed who had a confirmed or unconfirmed
complete remission after radiotherapy: 15 of 253
patients (6%) in the group received six cycles of
MOPP-ABV plus IF-RT, 14 of 259 patients (5%)
in the group received four cycles of MOPP-ABV
plus IF-RT, and 13 of 254 patients (5%) in the
group received MOPP-ABV plus subtotal nodal
RT [16]. There were no significant differences in
the 5-year event-free survival estimates among
the three groups.
The GHSG used a two-by-two factorial
design in the HD11 trial aimed at comparing
unfavorable early-stage HL using two different
chemotherapy regimen: 4xABVD vs. 4xBEACOPPbaseline (bleomycin, etoposide, adriamycin, cyclophosphamide, vincristin, procarbazin,
prednisone) as well as 30 Gy IF-RT vs. 20 Gy
[18]. Concerning RT, the 20 Gy arm was inferior
to 30 Gy when ABVD was used, but when BEACOPP was used, this difference disappeared and
20 Gy was equivalent to 30 Gy.
Taken together, 4xABVD and 30 Gy IF-RT
were considered as standard of care for early
unfavorable HL.
12.3.2 Chemotherapy
Besides the objective of reducing long-term toxicity with dose and field reductions, investigators
aimed at improving disease control further by
modifying chemotherapy schemes.
M. P. E. André and A. Engert
242
The GHSG HD11 study was based on a two-­
by-­two factorial design with the aim of comparing patients between two different regimen in
unfavorable early-stage HL: 4xABVD vs.
4xBEACOPPbaseline (bleomycin, etoposide,
adriamycin, cyclophosphamide, vincristin, procarbazin, prednisone) and 30 Gy IF-RT vs. 20 Gy
[18]. No improvement was demonstrated using
four cycles of BEACOPPbaseline compared with
four cycles of ABVD.
Similarly, the EORTC/LYSA H9U [19] study
compared 6 cycles of ABVD and 30 Gy IF-RT
(standard arm) with 4xABVD and 30 Gy IF-RT
and 4xBEACOPPbaseline followed by 30 Gy
IF-RT. Results in the 4xABVD and IF-RT (5-year
EFS, 85.9%) and the 4xBEACOPPbaseline and
IF-RT (5-year EFS, 88.8%) were not inferior to
6xABVD and IF-RT (5-year EFS, 89.9%) differences of 4.0% (90% CI, −0.7% to 8.8%) and of
1.1% (90% CI, −3.5% to 5.6%), respectively.
The 5-year OS estimates were 94%, 93%, and
93%, respectively. Because four cycles of
BEACOPPbaseline were more toxic but equally
efficient than four cycles of ABVD, it was not
considered as a new standard.
In their HD14 follow-up trial, the GHSG compared four cycles of ABVD and 30 Gy IF-RT
with two BEACOPPesc plus two ABVD and
30 Gy IF-RT. With a total of 1528 patients
included, a significant PFS advantage for « 2+2 »
compared with 4xABVD was detected with a
5-year PFS difference of 6.2% (95.4% vs. 89.1%;
HR, 0.45; 95% CI, 0.3–0.69) [20]. The « 2+2 »
approach, however, is associated with more
hematologic toxicity, but no difference in long-­
term toxicity or OS has been documented so far.
A longer follow-up will be needed to assess
potential risks and long-term benefits with intensive upfront therapy in patients with early-stage
unfavorable disease.
12.3.3 Chemotherapy Alone
Based on randomized trials performed in
advanced Hodgkin lymphoma patients and the
risk of late complications after radiotherapy, the
question arose whether RT can also be omitted in
unfavorable early stages. A number of trials conducted had important limitations: some trials
included pediatric patients, all stages of disease
used divergent definitions of unfavorable prognostic features, or there was a lack of statistical
power to detect clinically relevant differences in
PFS between RT and no-RT arms. The NCIC/
ECOG study on early stages had 12-year overall
survival as primary endpoint; patients with bulky
disease were excluded from entry [21]. This
study showed a significant 11% survival benefit
for treatment with ABVD alone as compared to
ABVD+STNI, notwithstanding a significant 8%
advantage in PFS for those who received combined modality treatment. The remarkable conversion of an inferior PFS to a superior long-term
OS for the ABVD-alone treatment arm was
mainly due to an excess of late toxic deaths in the
combined modality treatment (CMT): 23 vs.
11 in the former. These deaths were mainly due
to second cancers and intercurrent disease.
Admittedly, STNI has become outdated, but the
results corroborate the difficulties in interpreting
different treatment approaches with divergent
short-term (control of disease) and long-term
(toxicity) effects.
12.4
PET-Adapted Treatment
Strategies
12.4.1 Interim PET
In the publication of Gallamini et al., 260 newly
diagnosed HL patients were consecutively enrolled
in order to evaluate the prognostic role of an interim
PET-CT (iPET). Most of the patients were advanced
HL, and the study showed that iPET overshadows
the prognostic value of the International Prognostic
Score and emerges as the single most important tool
for planning of risk-adapted treatment in advanced
HL [22]. A similar evaluation conducted in 257
stage I to IIA patients treated with chemotherapy
plus radiation therapy led to similar conclusions
showing that iPET was a strong prognostic factor
for both, progression free and OS [23].
The standardization of iPET is critical for the
appropriate incorporation of this imaging modal-
12
Treatment of Early Unfavorable Hodgkin Lymphoma
ity into routine clinical practice. For this purpose,
successive international interpretation criteria
have been proposed and are regularly updated
according to improvement of diagnosis, treatment, and follow-up modalities. The current recommendation is to use the 2014 Lugano
Classification for response assessment but also
for staging of HL [24]. The Deauville 5-point
scale criteria (D5PS) allow for more accurate
measurement of response by using a categorical
scoring system designed for the visual interpretation of PET-CT. This score is now well validated
and reproducible [25].
However, it should be emphasized that the
definition of PET-CT negativity to escalate or deescalate therapy has been highly variable between
studies changing with the evolution of interpretation criteria. The actual recommendation is to
classify PET-CT with a D5SP <4 as negative and
D5SP >3 as positive. This categorization is also
in agreement with the PET-CT results of the
phase III H10 trial in early-stage HL recently
reanalyzed using the D5PS criteria showing that
patients with an interim PET-CT having a D5SP
<4 have a prognosis similar to those with D5PS
of 1 or 2 [9].
Subsequently, several trials were launched
with the aim to evaluate early treatment adaptation according to iPET results after 2 or 3 cycles
of ABVD.
12.4.2 Clinical Trials
12.4.2.1 Rapid Study
In the UK RAPID trial, 602 patients with newly
diagnosed stage IA or stage IIA HL received 3
cycles of ABVD and then underwent iPET [26].
RAPID included both, favorable (2/3) and unfavorable (1/3) early-stage HL in the same trial
according to GHSG or EORTC/LYSA risk classification. Patients with negative iPET (Deauville
score of 1 or 2) were randomly assigned to
receive IF-RT or no further treatment; patients
with positive iPET (Deauville score 3–5) received
a fourth cycle of ABVD and RT. The 3-year progression-free survival rate was 94.6% (95% CI,
91.5–97.7) in the RT group and 90.8% (95% CI,
243
86.9–94.8) in the group receiving no further therapy, with an absolute risk difference of −3.8 percentage points (95% CI, −8.8–1.3). As the upper
confidence interval limit exceeded the predefined
non-inferiority margin of 7%, the study did not
show non-inferiority of the strategy of no further
treatment. Nevertheless, patients in this study
with early-stage HL and negative iPET findings
after three cycles of ABVD had a very good prognosis either with or without consolidation radiotherapy. The impact on overall survival and late
effects needs additional follow-up.
12.4.2.2 H10 Study
Actually, the only published study to evaluate an
iPET approach in the specific group of unfavorable patients is H10 [27]. Unfavorable patients
were defined as age ≥50 years, large mediastinal
mass (M/T ratio >0.35), elevated erythrocyte sedimentation rate (with B symptoms, ≥30 mm/h;
without B symptoms, ≥50 mm/h), and >3 nodal
areas. Patients with a negative iPET were randomized between 4xABVD followed by IN-RT
(n = 292) or 6 cycles of ABVD (n = 302). After a
median follow-up of 5.1 years, a total of 54 PFS
events have occurred: 16 patients experienced
relapsed disease and 6 died from causes not related
to HL in the ABVD + IN-RT arm. In contrast, 30
patients experienced relapse and 2 died from
causes not related to HL in the ABVD-only arm.
Intention-to-treat 5-year PFS rates were 92.1%
(95% CI, 88.0–94.8) and 89.6% (95% CI, 85.5–
92.6) in the ABVD + IN-RT and ABVD-­
only
arms, respectively, with HR 1.45 (95% CI, 0.8–
2.5) favoring ABVD + IN-RT. Non-inferiority
could not be demonstrated as the upper bound of
the 95% CI for the estimated HR (2.50) exceeded
the prespecified non-inferiority margin (2.10).
However, the difference for the 5-year PFS was
only 2.5% (95% CI: −6.6% −0.5%) fitting in the
range of the 10% prespecified non-inferiority margin. Therefore, in this group of unfavorable
patients, the benefit of combined modality treatment seems to be less clinically relevant than in
the favorable group.
In the 594 unfavorable patients, 30/302 developed relapse after chemotherapy alone vs. 16/292
after CMT. Relapses after chemotherapy alone
M. P. E. André and A. Engert
244
Management of iPETPositive Patients
In the RAPID [26] and HD17 trial, patients with a
positive iPET received the standard arm of treatment. So far, only the data from RAPID are published. Among the 571 patients enrolled in the
study having an iPET after 3 ABVD, 145 were
iPET positive (D5PS 3–5). So far, 127 of the 145
12.4.2.3 Other Studies
patients (87.6%) in the group with positive PET
In the 50604 phase 2 trial, patients with non-­bulky findings were alive without disease progression.
stage I/II disease with a negative iPET after There had been 18 events in this group: 10 events
2xABVD (135 of 149 patients, Deauville score of disease progression (6.9% of the patients), 5
(DS), 1–3) were treated with an additional deaths with disease progression (3.4% of the
2xABVD without consolidative RT, whereas patients), and 3 deaths without disease progrespatients with a positive iPET (14 of 149 patients) sion (2.1% of the patients). A total of 8 of the 14
received 2xBEACOPPesc and 30 Gy patients (57.1%) in this group who required secIF-RT. Estimated 3 years PFS rates of 91% and ond-line treatment received high-­dose chemother66%, respectively, for the iPET-negative and PET- apy followed by autologous transplant.
positive cohorts were reported (p = 0.011), HR
In the H10 study, iPET-positive patients from
3.84 (95% CI, 1.50–9.84) [28]. These data sug- both favorable and unfavorable groups were
gest that four cycles of ABVD result in durable included together, because of their presumed
remissions for the majority of patients with early- shared poor prognosis, in a randomized trial comstage non-bulky HL and negative iPET.
paring 3-4xABVD and RT vs. 2xABVD + 2xBEAThe GHSG HD17 study evaluating iPET-­ COPPesc and IN-RT. In the overall iPET-­positive
adapted treatment in unfavorable patients has group (n = 361) and a median follow-up of 4.5
completed recruitment, but results are pending. years, a total of 57 events for PFS occurred: 41 (36
The trial compares 2xBEACOPPesc + 2xABVD, relapses and 5 deaths not related to HL) in the
and RT vs. 2xBEACOPPesc + 2xABVD in iPET- ABVD + IN-RT arm and 16 (13 relapses and 3
negative patients. Major difference comparing deaths not related to HL) in the BEACOPPesc +
the different studies are reported in (Table 12.3). IN-RT arm. Intent-to-treat 5-year PFS rates were
77.4% (95% CI, 70.4–82.9) and 90.6% (95% CI,
84.7–94.3) in the ABVD + INRT and
Table 12.3 Comparison of RAPID, H10, and HD17 BEACOPPesc + IN-RT arms, respectively, with an
trials
HR of 0.42 (95% CI, 0.23–0.74; P = 0.002) in favor
H10
RAPID
HD17
of BEACOPPesc + IN-RT. The 5-year OS rates
PET baseline
95%
0%
0%
were 89.3% vs. 96.0% for ABVD + IN-RT and
Interim PET
2xABVD
3xABVD 2xABVD
BEACOPPesc + IN-RT, respectively, with an HR
PET review
75%
100%
100%
of 0.45 (95% CI, 0.19–1.07; P = 0.062) (Fig. 12.2).
Noninferiority 10%
7%
occurred <2 years in 27/30 patients and in 3
patients after 2 years. Relapses after CMT
occurred <2 years in 8/16 patients, and in 8
patients after 2 years. Relapses after chemotherapy occurred mostly in initially involved areas in
26/30. After CMT, relapses in involved areas
were observed in 9/16 patients (Table 12.3).
margin
Stage
I–IIB (bulky)
Radiotherapy
IN RT 30Gy
PET
interpretation
I–IIA
IF RT
30Gy
5-point
International
Harmonization scale
Project [38]
I–IIB
(bulky
without
RF)
IF RT
30Gy
5-point
Deauville
score
12.4.2.4
12.5
ESMO and NCCN
Recommendations
The recently published ESMO guidelines recommend for intermediate stage: 4xABVD or
2xBEACOPPesc + 2xABVD and 30 Gy IS-RT or
2xABVD and an iPET, if the iPET is negative: 2
12
Treatment of Early Unfavorable Hodgkin Lymphoma
245
b
100
90
80
70
60
50
40
30
20
10
Overall Survival (%)
Progression -Free Survival (%)
a
HR, 0.42 (95% Cl, 0.23 to 0.74); P = .002
0
1
2
3
4
5
6
7
100
90
80
70
60
50
40
30
20
10
HR, 0.45 (95% Cl, 0.19 to 1.07); P = .062
0
8
1
2
O
41
n
3
4
5
6
7
8
Time (years)
Time (years)
No. at risk
O
192 167
156
147
105
57
21
0
ABVD + INRT
16 169 157
152
141
95
61
14
1
BEACOPPesc + INRT
n
No. at risk
18
192 189
181
167
119
65
24
0
ABVD + INRT
7
169 166
161
149
101
63
15
1
BEACOPPesc + INRT
Fig. 12.2 PFS (a) and OS (b) of patients iPET positive
patients included in the H10 trial. After 2xABVD patients
were randomized between 2xABVD and 30 Gy INRT vs.
2xBEACOPPesc and 30 Gy INRT. Reprinted from André
et al. with permission
additional ABVD and 30 Gy IS-RT and if iPET is
positive: 2 additional BEACOPPesc and 30 Gy
IS-RT [4].
The 2017 NCCN guidelines recommend a
PET-guided approach. For intermediate or unfavorable disease and bulky mediastinum, several
options are discussed including the HD14, H10
approaches but also Stanford V [29].
a randomized multicenter, phase II trial in order to
improve the PET response rate after 2 cycles with
BV-AVD for previously untreated, early-stage
unfavorable HL [31]. In total, 170 patients were
included, 113 were randomized in the BV-AVD
arm and 57 in the ABVD arm. After 2 cycles of
treatment, 93/113 patients (82.3%, 95% CI 75.3–
88.0) and 43/57 (75.4%, 95% CI 64.3–84.5)
achieved a negative PET (Deauville score 1–3)
based on central review in the experimental and
standard arms, respectively. With the lower bound
of the 90% confidence interval superior to 75% in
the experimental arm, the primary objective can be
considered to be met. An increased toxicity with
BV-AVD regimen compared to ABVD was
observed with a higher rate of grade 3–4 AEs and
SAEs during treatment. In another trial, Kumar
et al. treated 29 early-stage unfavorable patients
with 4 cycles of BV-AVD followed by 20 Gy
involved-site radiotherapy, and 90% of patients
achieved a negative PET after two cycles [32].
12.6
New Drugs
12.6.1 Brentuximab Vedotin
Brentuximab vedotin (BV) is an antibody–drug
conjugate composed of a CD30-targeted chimeric
monoclonal antibody covalently linked to the
microtubule disrupting agent monomethyl
auristatin E via a protease-cleavable linker. In a
phase 2 single-arm study, patients with relapsed or
refractory HL treated with BV after failure of
high-dose chemotherapy and post-autologous
stem-cell transplant, 76 (75%) of 102 patients
achieved an objective response, and 35 patients
(34%) achieved complete remission. Adverse
events were manageable with dose reduction or
delay. BV was also tested in combination with
AVD chemotherapy (BV-AVD) demonstrating
promising efficacy with a favorable safety profile
in a phase I trial for treatment-naive patients [30].
Based on these results, Fornecker et al. conducted
12.6.2 Checkpoint Inhibitors
Nivolumab and Pembrolizumab are immune
checkpoint inhibitors targeting the programmed
death-1 receptor [33, 34]. These checkpoint
inhibitors augment T-cell activation and restore
antitumor T-cell function. In the phase 2
CheckMate 205 study, nivolumab demonstrated
M. P. E. André and A. Engert
246
frequent (65–73%) and durable objective
responses across 3 cohorts of patients with
relapsed/refractory HL after failure of autologous
hematopoietic cell transplantation. Cohort D of
CheckMate 205 enrolled untreated patients with
advanced-stage newly diagnosed HL (stage III,
IV, or II with B symptoms and extranodal or
bulky disease). Nivolumab monotherapy followed by N-AVD combination therapy was well-­
tolerated and active in patients with newly
diagnosed, untreated, advanced-stage HL. This
combination of nivolumab and AVD is actually
being evaluated in phase II in early-stage unfavorable HL (NIVAHL, NCT03004833).
Pembrolizumab is also being evaluated in combination with AVD (NCT03226249).
12.7
Conclusions and Future
Strategies
The reduction of dose and size of RT and more
recently, PET-adapted strategies have reduced the
burden of treatment used to treat early-stage HL
and also defined a subpopulation that can benefit
from early intensification. Unfortunately, there is
no evidences that this could reduce long-term toxicities. Recently, three new drugs (brentuximab
vedotin, nivolumab, and pembrolizumab) showed
interesting results in the setting of relapsing
patients [35–37]. These drugs are now also being
actively evaluated in first-line for early-stage HL,
either alone (i.e., in elderly patients) or in combination with AVD. Phase II data are promising, but
only randomized phase III trial can change the
standard of care in this highly curable group of
patients. Finally, circulating cell-free DNA could
emerge as a very interesting tool to refine response
evaluation and better define cure [10].
References
1. Lister TA, Crowther D, Sutcliffe SB et al (1989)
Report of a committee convened to discuss the evaluation and staging of patients with Hodgkin’s disease:
Cotswolds meeting. J Clin Oncol 7(11):1630–1636
2. Kostakoglu L, Cheson BD (2014) Current role of
FDG PET/CT in lymphoma. Eur J Nucl Med Mol
Imaging 41(5):1004–1027
3. Stevens WB, van Krieken JH, Mus RD et al (2012)
Centralised multidisciplinary re-evaluation of diagnostic procedures in patients with newly diagnosed
Hodgkin lymphoma. Ann Oncol 23(10):2676–2681
4. Eichenauer DA, Aleman BMP, Andre M et al (2018)
Hodgkin lymphoma: ESMO Clinical Practice
Guidelines for diagnosis, treatment and follow-up.
Ann Oncol 29(Suppl 4):iv19–iv29
5. El-Galaly TC, d’Amore F, Mylam KJ et al (2012)
Routine bone marrow biopsy has little or no therapeutic consequence for positron emission tomography/computed tomography-staged treatment-naive
patients with Hodgkin lymphoma. J Clin Oncol
30(36):4508–4514
6. Klimm B, Goergen H, Fuchs M et al (2013) Impact
of risk factors on outcomes in early-stage Hodgkin’s
lymphoma: an analysis of international staging definitions. Ann Oncol 24(12):3070–3076
7. Specht L, Nordentoft AM, Cold S, Clausen NT,
Nissen NI (1988) Tumor burden as the most important
prognostic factor in early stage Hodgkin’s disease.
Relations to other prognostic factors and implications
for choice of treatment. Cancer 61(8):1719–1727
8. Gobbi PG, Ghirardelli ML, Solcia M et al (2001)
Image-aided estimate of tumor burden in Hodgkin’s
disease: evidence of its primary prognostic importance. J Clin Oncol 19(5):1388–1394
9. Cottereau AS, Versari A, Loft A et al (2018)
Prognostic value of baseline metabolic tumor volume
in e­ arly-­stage Hodgkin lymphoma in the standard arm
of the H10 trial. Blood 131(13):1456–1463
10. Spina V, Bruscaggin A, Cuccaro A et al (2018)
Circulating tumor DNA reveals genetics, clonal evolution, and residual disease in classical Hodgkin lymphoma. Blood 131(22):2413–2425
11. Schaapveld M, Aleman BM, van Eggermond AM
et al (2015) Second cancer risk up to 40 years after
treatment for Hodgkin’s lymphoma. N Engl J Med
373(26):2499–2511
12. Baxstrom K, Peterson BA, Lee C, Vogel RI, Blaes AH
(2018) A pilot investigation on impact of participation in a long-term follow-up clinic (LTFU) on breast
cancer and cardiovascular screening among women
who received chest radiation for Hodgkin lymphoma.
Support Care Cancer 26(7):2361–2368
13. Cutter DJ, Schaapveld M, Darby SC et al (2015) Risk
of valvular heart disease after treatment for Hodgkin
lymphoma. J Natl Cancer Inst 107(4):djv008
14. Hahn E, Jiang H, Ng A et al (2017) Late cardiac toxicity after mediastinal radiation therapy for Hodgkin
lymphoma: contributions of coronary artery and
whole heart dose-volume variables to risk prediction.
Int J Radiat Oncol Biol Phys 98(5):1116–1123
15. Engert A, Schiller P, Josting A et al (2003) Involved-­
field radiotherapy is equally effective and less toxic
compared with extended-field radiotherapy after four
cycles of chemotherapy in patients with early-stage
unfavorable Hodgkin’s lymphoma: results of the HD8
trial of the German Hodgkin’s Lymphoma Study
Group. J Clin Oncol 21(19):3601–3608
12
Treatment of Early Unfavorable Hodgkin Lymphoma
16. Ferme C, Eghbali H, Meerwaldt JH et al (2007)
Chemotherapy plus involved-field radiation in
early-stage Hodgkin’s disease. N Engl J Med
357(19):1916–1927
17. Sasse S, Brockelmann PJ, Goergen H et al (2017)
Long-term follow-up of contemporary treatment in
early-stage Hodgkin lymphoma: updated analyses of
the German Hodgkin Study Group HD7, HD8, HD10,
and HD11 Trials. J Clin Oncol 35(18):1999–2007
18. Eich HT, Diehl V, Gorgen H et al (2010) Intensified
chemotherapy and dose-reduced involved-field
radiotherapy in patients with early unfavorable
Hodgkin’s lymphoma: final analysis of the German
Hodgkin Study Group HD11 trial. J Clin Oncol
28(27):4199–4206
19. Ferme C, Thomas J, Brice P et al (2017) ABVD or
BEACOPPbaseline along with involved-field radiotherapy in early-stage Hodgkin Lymphoma with risk
factors: Results of the European Organisation for
Research and Treatment of Cancer (EORTC)-Groupe
d’Etude des Lymphomes de l'Adulte (GELA) H9-U
intergroup randomised trial. Eur J Cancer 81:45–55
20. von Tresckow B, Plutschow A, Fuchs M et al (2012)
Dose-intensification in early unfavorable Hodgkin’s
lymphoma: final analysis of the German Hodgkin
Study Group HD14 trial. J Clin Oncol 30(9):907–913
21. Meyer RM, Gospodarowicz MK, Connors JM et al
(2012) ABVD alone versus radiation-based therapy
in limited-stage Hodgkin’s lymphoma. N Engl J Med
366(5):399–408
22. Gallamini A, Hutchings M, Rigacci L et al (2007)
Early interim 2-[18F]fluoro-2-deoxy-D-glucose positron emission tomography is prognostically superior
to international prognostic score in advanced-stage
Hodgkin’s lymphoma: a report from a joint Italian-­
Danish study. J Clin Oncol 25(24):3746–3752
23. Simontacchi G, Filippi AR, Ciammella P et al (2015)
Interim PET after two ABVD cycles in early-stage
Hodgkin lymphoma: outcomes following the continuation of chemotherapy plus radiotherapy. Int J Radiat
Oncol Biol Phys 92(5):1077–1083
24. Cheson BD, Fisher RI, Barrington SF et al (2014)
Recommendations for initial evaluation, staging, and
response assessment of Hodgkin and non-Hodgkin
lymphoma: the Lugano classification. J Clin Oncol
32(27):3059–3068
25. Barrington SF, Kirkwood AA, Franceschetto A et al
(2016) PET-CT for staging and early response: results
from the response-adapted therapy in advanced
Hodgkin lymphoma study. Blood 127(12):1531–1538
26. Radford J, Illidge T, Counsell N et al (2015)
Results of a trial of PET-directed therapy for
early-stage Hodgkin’s lymphoma. N Engl J Med
372(17):1598–1607
247
27. Andre MPE, Girinsky T, Federico M et al (2017) Early
positron emission tomography response-adapted
treatment in Stage I and II Hodgkin lymphoma: final
results of the randomized EORTC/LYSA/FIL H10
trial. J Clin Oncol 35(16):1786–1794
28. Straus DJ, Jung SH, Pitcher B et al (2018) CALGB
50604: risk-adapted treatment of nonbulky early-stage
Hodgkin lymphoma based on interim PET. Blood
132(10):1013–1021
29. Hoppe RT, Advani RH, Ai WZ et al (2017) Hodgkin
Lymphoma Version 1.2017, NCCN Clinical Practice
Guidelines in Oncology. J Natl Compr Canc Netw
15(5):608–638
30. Younes A, Connors JM, Park SI et al (2013)
Brentuximab vedotin combined with ABVD or AVD
for patients with newly diagnosed Hodgkin’s lymphoma: a phase 1, open-label, dose-escalation study.
Lancet Oncol 14(13):1348–1356
31. Fornecker L, Lazarovici J, Aurer I et al (2017) PET-­
based response after 2 cycles of Brentuximab Vedotin
in combination with AVD for first-line treatment of
unfavorable early-stage Hodgkin lymphoma: first
analysis of the primary endpoint of breach, a randomized phase II Trial of Lysa-FIL-EORTC intergroup.
Blood 130(Suppl 1):736
32. Kumar A, Casulo C, Advani RH et al (2017) A Pilot
Study of Brentuximab Vedotin and AVD chemotherapy followed By 20 Gy involved-site radiotherapy in
early stage, Unfavorable Risk Hodgkin Lymphoma.
Blood 130(Suppl. 1):734
33. Ramchandren R, Fanale MA, Rueda A et al (2017)
Nivolumab for newly diagnosed advanced-stage
Classical Hodgkin Lymphoma (cHL): results from
the phase 2 checkmate 205 study. Blood 130(Suppl
1):651
34. UK RAPID study: NCT 00003849.
35. Chen R, Zinzani PL, Fanale MA et al (2017) Phase II
study of the efficacy and safety of Pembrolizumab for
relapsed/refractory classic Hodgkin lymphoma. J Clin
Oncol 35(19):2125–2132
36. Younes A, Gopal AK, Smith SE et al (2012) Results
of a pivotal phase II study of brentuximab vedotin for
patients with relapsed or refractory Hodgkin’s lymphoma. J Clin Oncol 30(18):2183–2189
37. Younes A, Santoro A, Shipp M et al (2016) Nivolumab
for classical Hodgkin’s lymphoma after failure of both
autologous stem-cell transplantation and brentuximab
vedotin: a multicentre, multicohort, single-arm phase
2 trial. Lancet Oncol 17(9):1283–1294
38. Juweid ME, Stroobants S, Hoekstra OS et al (2007)
Use of positron emission tomography for response
assessment of lymphoma: consensus of the Imaging
Subcommittee of International Harmonization Project
in Lymphoma. J Clin Oncol 25(5):571–578
Treatment of Advanced-Stage
Hodgkin Lymphoma
13
Alexander Fosså, René-Olivier Casasnovas,
and Peter W. M. Johnson
Contents
13.1
Introduction and Early History of Combination Chemotherapy 249
13.2
13.2.1
13.2.2
Fourth-Generation Regimens Hybrid and Alternating Regimens BEACOPP Escalated 251
251
251
13.3
ABVD or BEACOPP Escalated as Standard First-Line Treatment? 253
13.4
13.4.1
13.4.2
Outcome Prediction The International Prognostic Score Positron Emission Tomography 255
255
255
13.5
13.5.1
13.5.2
Response-Adapted Therapy De-escalation of Therapy in Early Responders Escalation of Therapy in Early Nonresponders 256
256
257
13.6
Introduction of Brentuximab Vedotin into First-Line Treatment 259
13.7
I ntroducing Programmed-­Death-­1 Inhibitors into First-Line
Treatment 260
13.8
The Role of Radiotherapy 260
13.9
Summary 261
References A. Fosså
Department of Oncology and National Resource
Centre for Late Effects After Cancer Therapy, Oslo
University Hospital, Oslo, Norway
KG Jebsen Centre for B-cell Malignancies,
University of Oslo, Oslo, Norway
R.-O. Casasnovas
Department of Haematology, University Hospital F.
Mitterrand and INSERM1231, Dijon, France
P. W. M. Johnson (*)
Cancer Research UK Centre, University of Southampton,
Southampton, UK
e-mail: johnsonp@soton.ac.uk
261
13.1
Introduction and Early
History of Combination
Chemotherapy
The definition of advanced stage in Hodgkin
lymphoma (HL) has evolved over time.
Megavoltage radiotherapy (RT) techniques
proved efficacious in stage I and IIA disease in
the 1950s and 1960s, whereas only few patients
with stage III disease were cured by RT alone,
despite extended fields of treatment [1]. At the
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_13
249
A. Fosså et al.
250
time, more than 95% of patients with stage IV
HL succumbed to their disease within 5 years.
The first encouraging trials on combination
chemotherapy included patients with stages III
and IV. Thus, from a historical perspective,
advanced disease may be defined by risk groups
where curative combination chemotherapy is
planned as the primary treatment, while RT is
either not used or not the major component.
The definition of advanced disease has therefore varied between academic study groups and
between trials. This should be borne in mind
when evaluating treatment recommendations
and results from trials. For most studies, stages
IIB, III, and IV are considered advanced disease, while stage IIA is sometimes included,
but only in the presence of other risk factors
such as bulky disease or multiple sites of
involvement.
With the advent of more exact diagnostic tools
and targeted treatment, the distinction between
classical HL (cHL) and nodular lymphocyte-­
predominant HL has become more robust and
more important for treatment. The current chapter covers the development of treatment for cHL
only, recognizing that in the past this distinction
was not as clear as it is today.
The introduction by DeVita and coworkers of
the MOPP regimen to treat patients with advanced
HL was a milestone in oncology [2, 3]. MOPP
resulted in long-term remission in nearly 50% of
patients with stage III and IV disease and has
been used for than 40 years. Bonadonna and
coworkers were the first to report on the importance of anthracyclines in developing ABVD [4].
ABVD replaced MOPP as the preferred first-line
treatment worldwide as the result of a series of
randomized trials comparing ABVD, MOPP,
and/or MOPP/ABVD alternating, sequential, or
hybrid regimens [5–10]. The results were better
for ABVD or ABVD-containing regimens than
for MOPP alone, with failure-free (FFS) or event-­
free survival (EFS) rates at 5–10 years of 50–80%
for ABVD-containing regimens, compared to
35.9% for MOPP in Bonadonna’s original trial
and 50% in the Cancer and Leukemia Group B
(CALGB) trial reported by Canellos in 1992
(Table 13.1).
The acceptance of ABVD over MOPP and
MOPP-like regimens during the 1980s and 1990s
was not only motivated by its greater efficacy but
also by concerns about toxicity. Follow-up of the
early trials showed that irreversible gonadal dysfunction as well as acute leukemia occurred far
Table 13.1 MOPP and ABVD in randomized trials
Trial
Bonadonna et al. [5]
Santoro et al. [6]
US Intergroup [7]
Viviani et al. [8]
Connors et al. [9]
Year
1986
Combination regimen
MOPP/ABVD
alternating
MOPP
Number of
patients
43
45
1987
3×MOPP-RT-3×MOPP
114
118
2003
3×ABVD-RT-­
3×ABVD
ABVD (6 cycles)
MOPP/ABV hybrid (6
cycles)
MOPP/ABVD
alternating (6 cycles)
MOPP/ABVD hybrid
(6 cycles)
MOPP/ABVD hybrid
(8 cycles)
MOPP/ABVD
alternating. (8 cycles)
419
1996
1997
433
211
204
153
148
Outcome
64.6% (FFP)
83.9% (OS)
35.9% (FFP)
63.9% (OS)
62.8% (FFP)
77.4% (OS)
80.8% (FFP)
67.9% (OS)
63% (FFS)
82% (OS)
66% (FFS)
81% (OS)
67% (FFP)
74% (OS)
69% (FFP)
72% (OS)
71% (FFS)
81% (OS)
67% (FFS)
83% (OS)
Follow-up and comments
8 years
7 years; (sub)total nodal
irradiation in all patients
5 years; MDS and AML
only in MOPP-treated
patients
10 years; stage IB and
IIA included
5 years; radiotherapy or
prolonged chemotherapy
after cycle 6 for PR
(continued)
13
Treatment of Advanced-Stage Hodgkin Lymphoma
251
Table 13.1 (continued)
Trial
CALGB [10]
Year
1992
Combination regimen
ABVD
Number of
patients
115
MOPP
123
MOPP/ABVD
alternating
123
Outcome
61% (FFS)
73% (OS)
50% (FFS)
66% (OS)
65% (FFS)
75% (OS)
Follow-up and comments
5 years
Abbreviations: CALGB Cancer and Leukemia Group B, GHSG German Hodgkin Study Group, FFS failure-free survival, FFP freedom from progression, FFTF freedom from treatment failure, OS overall survival, MDS myelodysplastic
syndrome, AML acute myelogenous leukemia
more often in patients treated with MOPP and
MOPP-like regimens. With improved control of
the lymphoma, there was an increasing need to
balance the likelihood of cure and the risk of serious or fatal complications from treatment. This
balance has since been a key concept in the development of better treatment options for advanced-­
stage cHL.
ABVD is a safe outpatient regimen without
the need for close white blood cell monitoring, is
feasible in most adult patients up to the age of 60,
and can be administered in less-developed healthcare systems [11]. However, bleomycin may
cause fatal lung toxicity, especially in older
patients, and a long-term higher risk of cardiac
morbidity has been reported for patients treated
with 6–8 cycles of ABVD [12, 13]. Furthermore,
Canellos and coworkers reported the long-term
outcome for 123 patients treated with ABVD for
advanced cHL in the CALGB trial with a FFS of
only 47% and overall survival (OS) 59% after 14
years [14]. Therefore, alternative approaches
were developed to improve these results.
consolidation RT as part of first-line treatment
in a varying proportion of patients (Table 13.2).
None of these studies has established any regimen as superior to ABVD, with the notable
exception of the BEACOPP escalated regimen
developed by the German Hodgkin Study Group
(GHSG).
13.2.2 BEACOPP Escalated
The development of BEACOPP was motivated by
the recognition that dose intensity plays a role in
chemotherapy of advanced HL. Hasenclever and
coworkers developed a statistical model of doseresponse characteristics for drugs used in 706
patients treated with COPP (cyclophosphamide,
vincristine, procarbazine, prednisone)/ABVDlike regimens [23]. The model was then used to
simulate the effect of dose escalation and changes
of schedule and architecture of COPP-­ABVD and
to design the BEACOPP regimen. G-CSF was
mandatory to compensate for the myelotoxic
effects. In a phase II study, the optimal doses of
the BEACOPP baseline and BEACOPP escalated
regimens were determined [24]. The subsequent
13.2 Fourth-Generation
HD9 trial of the GHSG found the predicted doseRegimens
response curve to be correct when comparing the
original COPP/ABVD to BEACOPP baseline and
13.2.1 Hybrid and Alternating
Regimens
BEACOPP escalated [20]. Results from 1195 randomized patients showed a clear superiority of 8
Several academic groups developed combina- cycles of BEACOPP escalated over BEACOPP
tion regimens containing different drugs with baseline and COPP/ABVD at 5 years. Importantly,
known efficacy in HL, and many of these fourth-­ follow-­up data at 10 years confirmed these results:
generation regimens were tested in randomized with a median follow-up of 112 months, freedom
trials against ABVD [15–22]. According to from treatment failure (FFTF) and OS rates were
standards of the time, most of these trials used 64 and 75% in the COPP/ABVD group, 70 and
A. Fosså et al.
252
Table 13.2 Fourth generation trials
Trial
GHSG HD6 [15]
Year
2004
Intergroup Italy [16]
2005
Combination regimen
COPP/ABV/IMEP
hybrid (4 cycles)
COPP/ABVD
alternating (4 cycles)
ABVD (6 cycles)
Stanford V (12 weeks)
UK Lymphoma
Group LY09 [17]
2005
MOPPEBVCAD
(6 cycles)
ABVD (6–8 cycles)
ChlVPP/EVA
(6–8 cycles)
UK and Italy [18]
2002
ChlVPP/PABlOE
(3–4 cycles
alternating)
ChlVPP/EVA hybrid
(6 cycles)
VAPEC-B (11 weeks)
UKNCRI [19]
2009
ABVD (6–8 cycles)
Stanford V
GHSG HD9 [20]
GHSG HD12 [21]
GHSG HD15 [22]
2003
2011
2012
COPP/ABVD
alternating
(8 cycles)
BEACOPP baseline
(8 cycles)
BEACOPP escalated
(8 cycles)
BEACOPP escalated
(8 cycles)
BEACOPP escalated
and baseline (4 + 4)
Number of
patients
Outcome
223
54% (FFTF)
73% (OS)
245
56% (FFTF)
73% (OS)
122
78% (FFS)
90% (OS)
107
54% (FFS)
82% (OS)
106
81% (FFS)
89% (OS)
394
75% (EFS)
79% (FFP)
90% (OS)
112
77% (EFS)
84% (FFP)
89% (OS)
282
74% (EFS)
78% (FFP)
87% (OS)
144
82% (FFP)
78% (EFS)
89% (OS)
138
62% (FFP)
58% (EFS)
79% (OS)
261
76% (PFS)
90% (OS)
259
74% (PFS)
92% (OS)
260
69% (FFTF)
83% (OS)
469
466
836
834
76% (FFTF)
88% (OS)
87% (FFTF)
91% (OS)
86.4%(FFTF)
92% (OS)
84.8% (FFTF)
90.3% (OS)
Follow-up and comments
7 years
5 years; radiotherapy to
initial bulk and residual
mass
3 years; stages I and II
included
5 years; radiotherapy to
initial bulk and residual
mass
5 years; radiotherapy to
initial bulk and splenic
deposits; stage I and II
disease included
5 years; radiotherapy to
initial bulk and residual
mass
5 years; second
randomization to
radiotherapy (initial bulk or
residual disease) versus
observation
5 years; radiotherapy to
PET positive residuals only
BEACOPP escalated
705
84.4% (FFTF);
(8 cycles)
91.9% (OS)
BEACOPP escalated
711
89% (FFTF)
(6 cycles)
95.3% (OS)
BEACOPP-14
710
85.4% (FFTF)
(8 cycles)
94.5%/(OS)
Abbreviations: UKNCRI United Kingdom National Cancer Research Institute Lymphoma Group, GHSG German
Hodgkin Study Group, FFS failure-free survival, FFP freedom from progression, FFTF freedom from treatment failure, EFS event-free survival, PFS progression-free survival, OS overall survival
13
Treatment of Advanced-Stage Hodgkin Lymphoma
80% in the BEACOPP baseline group, and 82 and
86% in the BEACOPP escalated group, respectively [25].
However, BEACOPP escalated is associated
with higher rates of acute and long-term toxicity.
The high rate of secondary malignancies, including myelodysplastic syndrome (MDS) or acute
myelogenous leukemia (AML) and solid cancers,
and high rates of infertility could be attributed to
the higher doses of alkylating agents and the frequent use of RT in the HD9 trial [20, 25]. The
subsequent GHSG HD12 trial aimed at de-­
escalating treatment by comparing four courses of
BEACOPP escalated with four courses of escalated and four courses of baseline BEACOPP
(“4 + 4”) [21]. The role of RT was tested by a second randomization of consolidating radiation to
initial bulky and residual disease versus no RT. At
5 years, OS was 91%, FFTF 85.5%, and
progression-­free survival (PFS) 86.2% with no
statistical difference between 8 cycles of
BEACOPP escalated and the 4 + 4 arm but no
apparent benefit in terms of toxicity with the
reduced regimen. Because a number of patients
randomized to the arm without RT were in fact
irradiated based on recommendations by a blinded
expert panel, the conclusions to be drawn from
this part of the trial are limited. Since there was no
relevant benefit in terms of toxicity in the 4 + 4
treated patients, 8 cycles of BEACOPP escalated
remained standard for advanced-stage HL patients
in the GHSG. Recent long-term updates of both
the HD9 and 12 trials confirm the superior initial
results for BEACOPP escalated [26].
The subsequent HD15 study also tested de-­
escalation of chemotherapy with a reduction in
the number of cycles from 8 to 6 and the introduction of a dose-dense BEACOPP baseline regimen (BEACOPP-14) [22]. In addition,
PET-guided omission of RT to residual disease
was investigated. A total of 2182 patients were
randomized among the 3 study arms. Surprisingly,
when comparing 6 cycles of BEACOPP escalated with 8 cycles, both PFS (90.3% versus
85.6%) and OS (95.3% versus 91.9%) were significantly better with the reduced number of
cycles. With omission of RT in cases of PET-­
253
negative residual masses, only 11% of all patients
received additional RT without compromising
tumor control, and the negative predictive value
for the end of treatment PET at 12 months was
94.1% [27]. In summary, HD15 established six
cycles of BEACOPP escalated as a new standard
of care to be tested further in PET response-­
adapted trials (see below).
With extensive evidence from clinical trials,
BEACOPP escalated is a highly effective regimen
for patients with advanced HL. Acute toxicity
requires monitoring of patients between cycles,
hospitalization in around one third of patients, and
vigilance for potentially lethal neutropenic infections. Special attention should be paid to patients
above the age of 50 years and patients of poor performance status at start of treatment. With the
reduction of cycle numbers and less use of consolidation RT, rates of secondary malignancies
seem to be decreasing, but determining the true
rate of such complications will need longer observation [26]. Infertility in both women and men
treated with BEACOPP is still a concern [28].
13.3
BVD or BEACOPP Escalated
A
as Standard First-Line
Treatment?
These developments led to the emergence of two
alternative strategies for the treatment of
advanced HL: balancing cure rates and toxicity,
the first strategy proposed ABVD as the standard
front line regimen, with salvage treatment, high-­
dose chemotherapy (HDCT), and autologous
stem cell transplantation (ASCT) for those
patients failing initial therapy. With this strategy,
the majority of patients could be cured with
ABVD without exposing them to the toxicity of
BEACOPP. The second strategy used BEACOPP
escalated as first-line treatment, aiming to cure as
many patients as possible with first-line therapy,
but accepting more toxicity for those patients
who might have been cured with a less intensive
approach. Data are now available from direct
comparative trails and from meta-analyses to
address this question.
A. Fosså et al.
254
Table 13.3 ABVD versus BEACOPP in direct comparisons
Combination
Number of
Study
Year
regimen
patients
Disease control p
HD 2000 [29] 2009 ABVD
99
68% (PFS)
0.038*
BEACOPP
98
81%
(4 esc +2 bas)
CEC
98
78%
168
73% (FFFP)
0.004
Michelangelo- 2011 ABVD
GITIL-IIL
BEACOPP (4
163
85%
[30]
esc + 4 bas)
91%
84%
89%
ns
Follow-up and
comments
5 years
7 years;
salvage
treatment
specified
4 years; EFS
primary
endpoint
ABVD (8
275
72.8% (PFS)
0.005 86.7% ns
cycles)
BEACOPP
274
83.4%
90.3%
(4 esc + 4 bas)
LYSA H34
2014 ABVD
77
75% (PFS)
0.007 92%
0.06 5 years
IPS 0–2 [32]
BEACOPP
68
93%
99%
(4 esc + 4 bas)
Abbreviations: GITIL Gruppo Italiano di Terapie Innovative nei Linfomi, IIL Intergruppo Italiano Linfomi, LYSA
Lymphoma Study Association EORTC European Organization for Research and Treatment of Cancer, IPS International
Prognostic Score, CEC cyclophosphamide, lomustine, vindesine, melphalan, prednisone, epidoxorubicin, vincristine,
procarbazine, vinblastine, bleomycin, FFFP freedom from first progression, EFS event-free survival, PFS progressionfree survival, p p-value for comparison, esc escalated, bas baseline
*
BEACOPP versus ABVD
EORTC
20012 IPS 3–7
[31]
2016
Overall
survival p
84%
ns
92%
Four studies have been conducted comparing
these two approaches, all smaller and so far with
shorter follow-up than the HD9 trial. All studies
compare four escalated BEACOPP followed by
two or four baseline BEACOPP with six to eight
cycles of ABVD (Table 13.3).
The Italian HD2000 trial enrolled 307 patients
in 3 different treatment arms showing a significant superiority of BEACOPP (4+2 fashion) over
6 cycles of ABVD in terms of PFS but not for OS
[29]. At 5 years, the PFS rate was 68% for ABVD
and 81% for BEACOPP: OS was 84% for ABVD
and 92% for BEACOPP, respectively. The data of
this trial were updated at 10 years follow-up, and
the authors were not able to confirm the superiority of BEACOPP over ABVD in terms of PFS,
mainly because of higher mortality from second
malignancies observed after BEACOPP [33].
The Michelangelo Foundation, the Gruppo
Italiano di Terapie Innovative nei Linfomi (GITIL),
and the Intergruppo Italiano Linfomi (IIL) study
compared 6–8 courses of ABVD and BEACOPP
given in 4 + 4 fashion plus preplanned salvage with
HDCT [30]. Patients with a higher risk profile
based on international prognostic score (IPS) of 3
and more were included. Two thirds of the patients
also received RT. The final analysis showed a
7-year rate of freedom from first progression
(FFFP) of 85% in patients who received initial
treatment with BEACOPP and 73% for those who
received ABVD (p = 0.004). A total of 65 patients
(20 in the BEACOPP group and 45 in the ABVD
group) needed HDCT. After completion of the
planned treatment including salvage therapy, the
7-year OS rates were 89 and 84%, respectively
(p = 0.39). This trial was not powered to detect differences in OS, and the conclusion on overall
treatment outcome should be cautioned.
The slightly larger intergroup trial organized
by the EORTC had a similar design as the
Michelangelo-GITIL-IIL study [31]; 8 courses of
ABVD were compared to BEACOPP 4 + 4 with
no RT allowed. With a median follow-up of 3.6
years at the time of the final report, the 4-year
rates for EFS and OS were similar in ABVD-­
treated (63.7% and 86.7%, respectively) and
BEACOPP-treated patients (69.3% and 91.5%).
However, the secondary endpoint, PFS, was significantly lower in the ABVD than in the
BEACOPP arm (72.8% versus 83.4%,
p = 0.0052). There were no clear differences in
toxicity between the two arms.
Patients with low-risk advanced-stage disease
(IPS 0–2) were enrolled in the H34 trial conducted by the Lymphoma Study Association
(LYSA) [32]. With 150 patients randomized in
13
Treatment of Advanced-Stage Hodgkin Lymphoma
255
this trial, the complete remission rate was 85%
for ABVD and 90% for BEACOPP. With a
median follow-up of 5.5 years, 7 patients died: 6
treated with ABVD and 1 with BEACOPP. The
PFS at 5 years was 75% and 93% (p = 0.007) and
the OS 92% and 99% (p = 0.06) for ABVD- and
BEACOPP-treated patients. Although the number of patients recruited in this trial was rather
small, these results suggest that BEACOPP is
also more effective than ABVD in advanced-­
stage patients with lower risk.
All trials comparing ABVD and BEACOPP
directly are smaller than the GHSG trials and evaluated different numbers of BEACOPP escalated
(4 + 4 or 4 + 2, escalated, and baseline, respectively). 6 cycles of escalated BEACOPP have been
shown by the GHSG to represent the most effective strategy. Since there was uncertainty regarding the difference in OS between ABVD and
BEACOPP, a network meta-analysis was performed to indirectly compare these and other regimens [34]. The final analysis included nearly
10,000 patients from 14 different trials.
Reconstructed survival data suggested that 6
cycles of BEACOPP escalated have a 10% advantage over ABVD in terms of OS at 5 years (95%
confidence interval 3–15%), offering advancedstage HL patients the highest chance of cure.
Another more recent meta-analysis with data from
four of the trials mentioned above (and including
one trial comparing ABVD and BEACOPP in
early unfavorable disease) confirmed the superiority of BEACOPP escalated over ABVD in term of
PFS and OS [35]. There was a significantly
increased occurrence of MDS or AML in
BEACOPP-based strategies (relative risk 3.90,
p = 0.02), but not for second malignancies in total.
The risk of infertility could not be assessed by
these meta-analyses due to lack of data.
order to balance efficacy and toxicity. The
development of the IPS paralleled the quest for
more effective regimens than ABVD and was
aimed to further assess each patient’s risk of
treatment failure under ABVD- and ABVD-like
regimens [36].
The score was derived from 5141 patients who
had been treated with ABVD-like regimens with
or without RT. Seven factors had similar independent prognostic effects: serum albumin of less
than 4 g/dL, hemoglobin level of less than 10.5 g/
dL, male sex, age of 45 years or older, stage IV
disease (according to the Ann Arbor classification), leukocytosis (white cell count of at least
15 × 10−9 L), and lymphocytopenia (lymphocyte
count of less than 0.6 × 10−9 or less than 8% of
white-cell count). As outlined, several studies
evaluating BEACOPP escalated have selected
patients not only based on stage and B symptoms,
but also on a higher IPS. It is assumed that the
more intensive treatment will have the greatest
effect in high-risk patients. This is in part supported by the results of the HD9 trial, where the
absolute benefit in terms of improved OS in those
treated with BEACOPP escalated seems greater
in intermediate risk than in low-risk patients,
although not in the high-risk group [25]. The
LYSA H34 results in low-risk patients suggest a
significant 18% benefit in PFS rates and a borderline significant improvement in OS of 9%, both at
5 years, but this requires confirmation [32].
13.4
Outcome Prediction
13.4.1 T
he International Prognostic
Score
With the choice between ABVD and BEACOPP,
it would be preferable to treat each advanced
cHL patient according to their individual risk in
13.4.2 Positron Emission
Tomography
Assessment of the risk of treatment failure by
IPS has been partly displaced by early
response evaluation. To determine the optimal
amount of treatment needed, functional imaging in the form of FDG-PET has been developed to provide an early indication of
chemosensitivity in HL. PET is discussed in
detail elsewhere in this book (see Chap. 7). In
patients with mostly advanced disease, retrospective studies showed that the early PET
response (after two cycles of ABVD) overshadowed the prognostic value of the IPS and
thus could be an important tool for response-
A. Fosså et al.
256
adapted treatment planning in advanced HL
[37]. The utility of FDG-PET has been facilitated by development of a highly reproducible
5-point scale, the Deauville scale, for reporting
results [38].
Several recent studies have tested early
response evaluation by FDG-PET as a means of
adjusting subsequent therapy according to the
response to initial treatment. Strategies based on
initial ABVD or BEACOPP have helped to
define new standards of care in which each
patient receives as much therapy as deemed
necessary.
Table 13.4 Response adapted therapy in randomized trials
%PET-2
Upfront
Number of negative/
positive
Study
Year regimen
patients
RATHL
2016 ABVD
1203
86
[39]
DS 1-3
13.5
Response-Adapted Therapy
13.5.1 De-escalation of Therapy
in Early Responders
Groups using either ABVD or BEACOPP as initial therapy have aimed to reduce treatment intensity in patients with a negative interim PET, by
omitting potentially toxic components of the
regimens, by reducing the number of cycles, or
by omitting RT (Table 13.4).
In the international RATHL trial, all patients
initially received two cycles of ABVD [39].
Follow-up and
comments
3 years; ABVD vs
AVD in PET-2
negative disease
16
67.5% (BEACOPP)
87.8% BEACOPP
DS 4-5
escalated/-14 in
PET-2 positive
disease
2018 ABVD
782
81
87% (ABVD)
99%
3 years; with or
GITIL/
DS 1-3
without RT in bulky
FIL0607
PET-2 negative
[40]
disease
19
63% (R-BEACOPP) 89%
BEACOPP with or
DS 4-5
57% (BEACOPP)
without R in PET-2
positive disease
2016 BEACOPP 275
52
92.2%
97.7% 5 years; 6–8 vs 4
GHSG
esc
DS 1-2
(4×BEACOPPesc)
cycles
HD18
[41]
90.8%
95.4% BEACOPP
escalated in PET-2
(6–8×BEACOPPesc)
negative disease
48
88.4%
93.6% BEACOPP
DS 3-5
(R-BEACOPPesc)
escalated with or
96.4% without R in PET-2
89.7%
positive disease
(BEACOPPesc)
LYSA
2018 BEACOPP 823
87
89.4% (ABVD)
96.4% 5 years; modified
AHL2011
esc
DS 1-3
Deauville criteria
for PET-2, OS
13
88.4%
99%
similar in
DS 4-5
(BEACOPPesc)
PET-driven and
standard arm
Abbreviations: RATHL Response adapted therapy in Hodgkin lymphoma, GITIL/FIL Gruppo Italiano Terapie Innovative
Linfomi/Fondazione Italiana Linfomi, LYSA Lymphoma Study Association, GHSG German Hodgkin Study Group,
PFS progression-free survival, OS overall survival, DS Deauville score, PET-2 PET scan performed after 2 cycles of
therapy, BEACOPPesc BEACOPP escalated, R Rituximab
PFS
85% (ABVD)
84.4% (AVD)
OS
97.2%
97.6%
13
Treatment of Advanced-Stage Hodgkin Lymphoma
Using a Deauville score cut-off between 3 and 4,
83.7% of the patients were interim PET negative
and were randomized. In the experimental arm,
bleomycin was not given in the remaining four
cycles. The results showed 3-year PFS rates of
85.7% and 84.4%, respectively, in the standard
ABVD and AVD groups. Pulmonary toxicity was
reduced in patients treated with AVD in cycles 3
through 6. Thus, bleomycin is redundant in
advanced-stage patients who have achieved a
complete metabolic response after two cycles of
chemotherapy.
In patients with advanced HL who have a negative PET, defined as Deauville scores 1–3, after
two and six cycles of ABVD, the GITIL/
Fondazione Italiana Linfomi (FIL) HD0607 trial
demonstrated that consolidation RT to initially
bulky lesions can be omitted without a decrease
in tumor control at 3 years [40]. The 3-year PFS
was 87% in the interim PET-negative group.
The proportions of interim PET-negative
patients in the RATHL and GITL/FIL HD0607
studies and 3-year PFS rates are remarkably similar. They are comparable to other prospective
phase II trials using the same approach [42, 43].
However, the negative predictive value of interim
PET-CT in these prospective trials appears lower
than anticipated from previous retrospective
studies using non-standardized criteria for reporting PET results. This means that the largest number of treatment failures in advanced-stage HL
treated initially with ABVD will occur in the
interim PET-negative group.
Interim PET-CT has also been used to de-­
escalate therapy in early responders to BEACOPP
escalated. Using the FDG uptake in the mediastinal
blood pool as reference, corresponding to Deauville
scores of 1 and 2, the randomized GHSG HD18
trial demonstrated that the number of cycles of
BEACOPP could be reduced from six to four in
patients with a negative interim PET after two
cycles [41]. With 1005 out of 1945 randomized
patients (52%) having a negative interim PET by
these criteria, the HD18 trial indicated an excellent
5-year OS rate of 95% and a significant reduction of
severe acute hematological and non-­hematological
toxicities. The GHSG later provided post hoc
evidence that the excellent outcome of PET-
257
negative patients also holds for those with an interim
Deauville score of 3, and despite the fact that these
had all received six cycles of BEACOPP escalated
in the trail, a total of four escalated BEACOPPs to
patients with an interim score of 1–3 are currently
recommended [44].
The randomized LYSA AHL2011 trial evaluated whether a PET-driven strategy allows for a
tailored shift from BEACOPP escalated to
ABVD in advanced HL [45]. In the experimental
arm, patients with a negative PET after two
cycles of BEACOPP escalated completed treatment with four cycles of ABVD, while the
interim PET-­
positive patients continued with
four cycles of BEACOPP escalated. Patients
assigned to the standard arm received a total of
six cycles of BEACOPP escalated. After a
median observation time of 50 months, 5-year
PFS rates did not differ, 86.5% and 85.7% in the
standard and experimental arms, respectively.
Using a Deauville score cut-off between 3 and
4 in the experimental arm, 84% of patients
received 2 escalated BEACOPP and 4 ABVD
with a significant reduction in toxicity. Thus, it
appears possible to switch to ABVD on the basis
of a negative interim PET after two cycles of
escalated BEACOPP without loss of tumor control. By extrapolation from the RATHL study, it
may be possible to switch to AVD and thereby
avoid the risk of continued bleomycin exposure.
The LYSA AHL2011 used a modified Deauville
score for assessment of interim PET-CT defining
score 4 and 5 as a residual lesion uptake equal or
higher than 140% and 200% of the liver uptake,
respectively. These more stringent criteria may
explain the lower rate of interim PET-positive
patients in the AHL2011 (12.6%) than in the
HD18 trial (24%) and the higher treatment failure rate in the PET-positive group in AHL2011
compared to HD18 study.
13.5.2 E
scalation of Therapy in Early
Nonresponders
Several groups tested prospectively the approach
of escalating treatment in patients not responding
to two cycles of ABVD as defined by PET posi-
258
tivity (Table 13.4). With the limitations mentioned above, based on the historical data, these
patients have a very poor outcome with ABVD or
ABVD-like therapy. The 2- or 3-year PFS is
reported between 6% and 38% [46]. Most investigators have therefore considered it unethical in
these patients to compare any experimental therapy to ABVD in a randomized fashion.
The RATHL trial tested the escalation to either
BEACOPP escalated four cycles or BEACOPP-14
six cycles in interim PET-positive patients [39].
Of the 182 patients with positive findings on
interim PET-CT scans according to protocol, 94
received BEACOPP-14 and 78 received escalated BEACOPP. The results of a third PET-CT
scan were available for 160 patients, of whom
119 (74.4%) had negative findings. The 3-year
PFS rate for the group as a whole was 67.5%, and
the OS rate was 87.8%.
Similar results were reported in the GITIL/
FIL HD0607 trial [40]. 149 of 150 patients with
interim PET-positive results continued on
BEACOPP (4 escalated and 4 baseline) with or
without the addition of rituximab. After four
BEACOPP escalated, a PET evaluation was performed in 136 patients. At the time of the report,
disease progression was registered in 27 of 108
PET-negative scans compared with 25 of 28 PET-­
positive scans. The 3-year PFS rate was 60%, and
the 3-year OS rate was 89% in the interim PET-­
positive group.
The American South West Oncology Group
(SWOG) followed the same principles of escalating to six cycles of BEACOPP escalated after
ABVD with similar results for the interim PET-­
positive group compared to RATHL and GITIL/
FIL HD0607 trials [42]. Escalation to an alternative regimen consisting of ifosfamide, gemcitabine and vinorelbine (IGEV) followed by
HDCT was pursued in the Italian HD0801 study
[43]. In an intention-to-treat analysis, the authors
reported for the PET-positive patients (excluding
those with Deauville score 3) a 2-year PFS of
75%.
Taking together the shift from ABVD to
BEACOPP escalated or variants thereof seems
justified in interim PET-positive patients.
Whether other salvage regimens including HDCT
A. Fosså et al.
might improve the outcomes further is unknown
at present. Nonetheless, the results remain suboptimal, and further improvements, possibly by
incorporation of novel drugs, are needed.
Following the success with initial BEACOPP
escalated in the HD9, HD12, and HD15 trials, the
GHSG tested whether it might be possible to
improve the results by adding rituximab, a monoclonal antibody binding to CD-20 on the surface
of B cells, in patients with interim PET-positive
disease [47]. The rationale is derived from the
recognition of a minority of cHL cases with
CD-20-positive Hodgkin or Reed-Sternberg cells
and the possible benefit of targeting normal B
cell in the tumor microenvironment. In the HD18
trial, patients with interim PET-positive disease
(Deauville score 3–5) after two cycles of escalated BEACOPP were randomized to six further
cycles with or without rituximab. The PFS rates
at 3 years for PET-positive patients were 93.0%
and 91.4% with or without rituximab, respectively, better than expected from the previous
HD15 trial and without any benefit of adding
rituximab.
Patients with an interim Deauville score 4
(post hoc analysis of HD18) or 4–5 (LYSA
AHL2011) after 2 BEACOPP escalated have
reduced PFS rates compared to interim PET-­
negative patients with a score of 1–3. As mentioned, a later post hoc analysis of the HD18
study showed that in patients receiving 6 cycles
of BEACOPP escalated, 3-year PFS rates were
92.2%, 95.9%, and 87.6% with interim PET
scores of 1–2, 3, and 4, respectively [44]. The
univariate hazard ratio (HR) for PFS in patients
with score 4 versus score 1–3 was 2.3 (p = 0.002).
Deauville score of 4 was the only factor remaining significant for PFS in a multivariate analysis
including the associated baseline risk factors. In
the LYSA AHL 2011 trial, using more stringent
definitions of Deauville score 4 and 5, interim
PET positivity was associated to a higher risk of
relapse or progression with 5-year PFS of 70.7%
versus 88.9% and a HR = 3.59 (p < 0.0001).
Whether these results for interim PET-positive
patients after initial BEACOPP therapy can be
improved, i.e., by the incorporation of novel
drugs, has not been tested yet.
13
Treatment of Advanced-Stage Hodgkin Lymphoma
The concept of response-adapted therapy in
advanced-stage HL has considerably changed the
way in which patients are treated. Unfortunately,
there is no direct comparison of strategies starting with ABVD and escalating in poor responders versus starting with BEACOPP escalated and
reducing treatment in the early responders.
13.6
Introduction of Brentuximab
Vedotin into First-Line
Treatment
With the approval of brentuximab vedotin (BV)
for relapsed and refractory disease (see Chap.
21), a novel targeted drug has been introduced
into the treatment of cHL. This has shown an
excellent balance of efficacy and tolerability in
persistent or recurrent disease [48]. Therefore,
BV is an ideal candidate to improve both the
ABVD and the BEACOPP regimens.
BV was initially combined with ABVD in a
phase I study; however, life-threatening pulmonary toxicity in this bleomycin-containing combination was observed [49]. BV at a fixed dose
(1.2 mg/kg body weight) was given safely with
bleomycin-deleted AVD (given the name
A+AVD, as Adcetris, the proprietory name for
BV) to 26 patients. 25 of the 26 (96%) had complete response after treatment. Neuropathy, probably as a result of coadministration of two
microtubule-disrupting agents (monomethyl
auristatin E and vinblastine), neutropenia and
complications thereof, including the need to add
growth factors in a number of patients, were
noted as relevant toxicities. In a recent update on
long-term outcome, the 5-year FFS and OS were
79% and 92% after ABVD plus BV and 92% and
100% after A+AVD, respectively [50].
After the initial encouraging results with
A+AVD, this regimen has been tested in a large
company-sponsored trial, comparing the new regimen to ABVD [51]. Patients with stage III and IV
disease were entered and interim PET-CT after
two cycles guided an optional switch to alternative frontline therapy at the treating physician’s
discretion, but only for patients with a Deauville
score of 5. The primary endpoint was modified
259
PFS, defined as time to disease progression, death,
or modified progression (with the latter defined as
evidence of non-complete response after completion of frontline therapy according to review by an
independent committee, followed by subsequent
anticancer therapy). With 1334 randomized
patients and a median follow-up of 24.6 months,
the 2-year modified PFS rates in the A+AVD and
ABVD groups were 82.1% and 77.2%, respectively, a difference of 4.9 percentage points
(p = 0.04). There was higher rate of febrile neutropenia in the A+AVD group, and growth factor
support was introduced as mandatory during the
course of the trial. There was also more severe
neuropathy in the experimental arm, but as
reported for other BV trials, this appeared to be
reversible in most patients. As expected from the
omission of bleomycin, lung toxicity was reduced.
These results produced a debate as to whether
A+AVD should be accepted as a new standard of
care. Follow-up is short, and the new endpoint,
incorporating into the definition of progression
any treatment given to patients with residual PETpositive disease (Deauville scores 3–5), hampers
comparison with other trials. With a modest difference between treatment arms, increased toxicity, and the absence of a documented survival
benefit, the same principles as in the debate over
ABVD and BEACOPP will apply. Another consideration is the considerable drug cost of BV and
expense of the necessary growth factors.
An interesting aspect of the Echelon-1 trial is
the prospective less stringent approach to PET-­
guided adaptation, where patients with an interim
PET-CT of 4 or 5 were either planned to continue
treatment with A+AVD or ABVD (Deauville
score 4) or allowed to do so as an alternative to
intensive chemotherapy (Deauville score 5). The
data thus reflect the outcome of ABVD-treated
interim PET-positive patients in the modern era,
using PET-CT in a prospective way and with uniform criteria comparable to other studies [52].
PET positivity rates were strikingly lower than in
other similar studies, with Deauville ≥4 found in
7% (47/644) in the A+AVD arm and 9% (58/670)
with ABVD; 5 patients with a Deauville score of
5 switched to alternative frontline therapy.
Subgroup analyses of the ABVD arm showed a
A. Fosså et al.
260
2-year modified PFS rate of 80.9% versus 42.0%
for interim PET-negative and PET-positive
patients, respectively.
Following the aim to improve tolerability
while maintaining efficacy, the GHSG has modified the BEACOPP regimen and introduced
BV. A randomized phase II trial testing six cycles
of two variants of BEACOPP was published
recently: in one arm, vincristine was replaced by
BV and bleomycin omitted (BrECAPP) [53]. A
more experimental regimen additionally replaced
dacarbazine for procarbazine and short-term
dexamethasone instead of long-term prednisone
(BrECADD). 104 patients were enrolled to the
study (52 were assigned to each study arm).
Complete responses were seen at completion of
BrECAPP and BrECADD in 86% and 88% of
patients, respectively. Particularly, the BrECADD
regimen was associated with a more favorable
toxicity profile and therefore selected for comparison to standard BEACOPP escalated in
advanced cHL in the ongoing HD21 trial
(NCT02661503). This trial is testing 4/6 cycles
of escalated BEACOPP (in a PET response-­
adapted design) against 4/6 cycles of BrECADD.
13.7
Introducing Programmed-­
Death-­1 Inhibitors into
First-Line Treatment
The second class of drugs recently introduced for
relapsed and refractory cHL is the checkpoint
inhibitors targeting programmed-death receptor
(PD) 1, on the surface of cells in the microenvironment of cHL. It appears that Hodgkin and ReedSternberg cells by expression of the PD-ligand
(PD-L)-1 and PD-L2 orchestrate the tissue microenvironment for tumor survival and growth
(reviewed in Chap. 22). These mechanisms seem to
be operable also in previously untreated cases [54].
Studies are underway to explore the effect of PD-1
inhibition also in the context of first-line treatment
of advanced disease. So far, no results are reported.
An initial study using nivolumab as sole initial therapy prior to the addition of AVD showed a response
rate of 69% for monotherapy and promising short
term outcomes, with PFS of 92% at 9 months [55].
13.8
The Role of Radiotherapy
The use of RT in advanced-stage HL has evolved
over time, from subtotal or total nodal RT (i.e.,
treating also areas of initial possible microscopic
disease) to involved-field RT (i.e., treating only
areas of initial macroscopic disease) or as RT in
situations associated with a higher risk of local
relapse (most often site of initial bulky lesions or
residual macroscopic disease possibly representing active tumor). The role of consolidation RT
for advanced HL also depends on the efficacy of
the prior chemotherapy (see Chap. 9 for further
details).
After MOPP or MOPP-like regimens, there
appeared to be a potential advantage of IFRT in a
meta-analysis of 16 randomized studies, whereas
this advantage is not evident after ABVD or
ABVD-like regimens [56, 57]. A randomized
EORTC study demonstrated that consolidation
IFRT did not improve outcomes in patients in complete remission after six to eight courses of MOPPABV, but potentially improved the outcome of
patients with a partial response [58]. A randomized
Groupe d’Etude des Lymphomes de l’Adulte
(GELA) trial showed that consolidation with subtotal or total nodal RT for patients in remission
after doxorubicin-containing systemic treatment
was not superior to two additional cycles of chemotherapy [59]. Thus, patients achieving a complete
radiological response with ABVD or ABVD-like
regimens do not need consolidation RT.
In current treatment algorithms, the chemosensitivity and quality of remission in each individual patient are assessed by FDG-PET, either
as an interim PET during or a PET done at the
end of treatment. PET results may be especially
helpful in assessing the need for consolidation in
situations with residual fibrotic masses after chemotherapy. The GITIL/FIL HD0607 trial randomized patients with advanced HL with a large
nodal mass at diagnosis (≥5 cm) and a negative
PET after two and six cycles of ABVD to RT of
the site of the initial large mass or observation
alone [40]. In 296 randomized patients, there
was no significant PFS improvement at 3 years,
97% versus 93%, respectively (p = 0.29).
Together, patients treated with ABVD that are in
13
Treatment of Advanced-Stage Hodgkin Lymphoma
261
complete remission by radiological criteria or
PET negative after two cycles can safely omit
RT, even in presence of residual masses. Whether
certain patients with residual masses that are
PET negative at the end of treatment (in the
absence of any interim PET) or patients with
interim or end of treatment positive PET in the
context of ABVD benefit from the addition of
RT has not been analyzed systematically.
The HD15 trial prospectively analyzed the
omission of RT in cases of PET-negative residual
masses after six or eight cycles of BEACOPP
[27]. All patients with residual disease of ≥2.5 cm
after chemotherapy were evaluated using additional PET and based on the criteria at the time;
those with a PET-positive result were irradiated
to the site of residual disease. Only 11% of all
patients received additional RT without compromising tumor control, and the negative predictive
value for the end of treatment PET at 12 months
was 94.1%. PFS at 4 years was similar for patients
without residual disease and patients with PET-­
negative residuals, 92.1% and 92.6%, respectively, showing that RT can be safely omitted in
both situations. PFS for PET-negative or PET-­
positive patients was 92.6% and 86.2% at 48
months (p = 0.022). Thus, a positive PET after
chemotherapy was associated with higher risk of
subsequent treatment failure, even though PET-­
positive patients were treated with additional
RT. The frequency and pattern of relapses still
suggest that local RT to PET-positive residual
disease is sufficient for these patients [60].
omission of bleomycin and/or RT is possible in
interim PET-negative patients. In case of an inadequate response, a shift to BEACOPP escalated may
still cure a reasonable number of patients, subjecting only a minority of patients to the increased toxicity associated with BEACOPP. Similarly, the
total number of cycles can be reduced from six to
four or therapy switched to four ABVD in patients
with interim negative PET results after two cycles
of BEACOPP escalated. RT can be safely omitted
in BEACOPP-treated patients with an end of treatment negative PET. With still short follow-up, the
OS in either approach is excellent, with the interim
PET-positive group after ABVD representing a
candidate group for implementation of novel
approaches. Apart from these more personalized
treatment strategies, early results from a combination of BV with AVD show modest improvements
in disease control but increased toxicity and costs.
BV is also tested in the context of a modified
BEACOPP regimen to enhance tolerability. After
decades of substantial but slow advances in the
treatment of advanced-stage HL, personalized
treatment strategies have resulted in better treatment options for our patients. Targeted and immunological approaches may improve results further
in the near future.
13.9
Summary
Advanced-stage HL has become a curable disease
for the majority of patients, and treatment decisions need to take into account the risk of serious
long-term consequences of therapy. After a decade
of debate whether first-line treatment with six to
eight cycles of ABVD or BEACOPP escalated
best balances the likelihood of cure and risk of
complications, individualized treatment based on
early response to chemotherapy has now
become a standard in many developed countries. Starting with ABVD, de-escalation by
References
1. Kaplan HS (1972) Hodgkin’s disease. Harvard
University Press, Cambridge, MA. 452 p
2. Devita VT Jr, Serpick AA, Carbone PP (1970)
Combination chemotherapy in the treatment of
advanced Hodgkin’s disease. Ann Intern Med
73(6):881–895
3. DeVita VT Jr, Simon RM, Hubbard SM, Young RC,
Berard CW, Moxley JH 3rd et al (1980) Curability
of advanced Hodgkin’s disease with chemotherapy.
Long-term follow-up of MOPP-treated patients
at the National Cancer Institute. Ann Intern Med
92(5):587–595
4. Bonadonna G, Zucali R, Monfardini S, De Lena
M, Uslenghi C (1975) Combination chemotherapy
of Hodgkin’s disease with adriamycin, bleomycin, vinblastine, and imidazole carboxamide versus
MOPP. Cancer 36(1):252–259
5. Bonadonna G, Valagussa P, Santoro A (1986)
Alternating non-cross-resistant combination chemotherapy or MOPP in stage IV Hodgkin’s disease. A report of 8-year results. Ann Intern Med
104(6):739–746
262
6. Santoro A, Bonadonna G, Valagussa P, Zucali R,
Viviani S, Villani F et al (1987) Long-term results of
combined chemotherapy-radiotherapy approach in
Hodgkin’s disease: superiority of ABVD plus radiotherapy versus MOPP plus radiotherapy. J Clin Oncol
5(1):27–37
7. Duggan DB, Petroni GR, Johnson JL, Glick JH, Fisher
RI, Connors JM et al (2003) Randomized comparison
of ABVD and MOPP/ABV hybrid for the treatment of
advanced Hodgkin’s disease: report of an intergroup
trial. J Clin Oncol 21(4):607–614
8. Viviani S, Bonadonna G, Santoro A, Bonfante V,
Zanini M, Devizzi L et al (1996) Alternating versus
hybrid MOPP and ABVD combinations in advanced
Hodgkin’s disease: ten-year results. J Clin Oncol
14(5):1421–1430
9. Connors JM, Klimo P, Adams G, Burns BF, Cooper
I, Meyer RM et al (1997) Treatment of advanced
Hodgkin’s disease with chemotherapy--comparison of
MOPP/ABV hybrid regimen with alternating courses
of MOPP and ABVD: a report from the National
Cancer Institute of Canada clinical trials group. J Clin
Oncol 15(4):1638–1645
10. Canellos GP, Anderson JR, Propert KJ, Nissen N,
Cooper MR, Henderson ES et al (1992) Chemotherapy
of advanced Hodgkin’s disease with MOPP, ABVD,
or MOPP alternating with ABVD. New Engl J Med
327(21):1478–1484
11. Geel JA, Chirwa TC, Rowe B, Eyal KC, Omar F,
Stones DK et al (2017) Treatment outcomes of
­children with Hodgkin lymphoma between 2000 and
2010: First report by the South African Children’s
Cancer Study Group. Pediatr Blood Cancer 64:10
12. Myrehaug S, Pintilie M, Tsang R, Mackenzie R,
Crump M, Chen Z et al (2008) Cardiac morbidity
following modern treatment for Hodgkin lymphoma:
supra-additive cardiotoxicity of doxorubicin and radiation therapy. Leukemia Lymphoma 49(8):1486–1493
13. Evens AM, Hong F, Gordon LI, Fisher RI, Bartlett
NL, Connors JM et al (2013) The efficacy and tolerability of adriamycin, bleomycin, vinblastine, dacarbazine and Stanford V in older Hodgkin lymphoma
patients: a comprehensive analysis from the North
American intergroup trial E2496. Brit J Haematol
161(1):76–86
14. Canellos GP, Niedzwiecki D (2002) Long-term follow-­up of Hodgkin’s disease trial. New Engl J Med
346(18):1417–1418
15. Sieber M, Tesch H, Pfistner B, Rueffer U, Paulus
U, Munker R et al (2004) Treatment of advanced
Hodgkin’s disease with COPP/ABV/IMEP versus
COPP/ABVD and consolidating radiotherapy: final
results of the German Hodgkin’s Lymphoma Study
Group HD6 trial. Ann Oncol 15(2):276–282
16. Gobbi PG, Levis A, Chisesi T, Broglia C, Vitolo
U, Stelitano C et al (2005) ABVD versus modified
Stanford V versus MOPPEBVCAD with optional and
limited radiotherapy in intermediate- and advanced-­
stage Hodgkin’s lymphoma: final results of a multicenter randomized trial by the Intergruppo Italiano
Linfomi. J Clin Oncol 23(36):9198–9207
A. Fosså et al.
17. Johnson PW, Radford JA, Cullen MH, Sydes MR,
Walewski J, Jack AS et al (2005) Comparison of
ABVD and alternating or hybrid multidrug regimens for the treatment of advanced Hodgkin’s lymphoma: results of the United Kingdom Lymphoma
Group LY09 Trial (ISRCTN97144519). J Clin Oncol
23(36):9208–9218
18. Radford JA, Rohatiner AZ, Ryder WD, Deakin DP,
Barbui T, Lucie NP et al (2002) ChlVPP/EVA hybrid
versus the weekly VAPEC-B regimen for previously untreated Hodgkin’s disease. J Clin Oncol
20(13):2988–2994
19. Hoskin PJ, Lowry L, Horwich A, Jack A, Mead B,
Hancock BW et al (2009) Randomized comparison of
the Stanford V regimen and ABVD in the treatment
of advanced Hodgkin’s Lymphoma: United Kingdom
National Cancer Research Institute Lymphoma
Group Study ISRCTN 64141244. J Clin Oncol
27(32):5390–5396
20. Diehl V, Franklin J, Pfreundschuh M, Lathan B,
Paulus U, Hasenclever D et al (2003) Standard and
increased-dose BEACOPP chemotherapy compared
with COPP-ABVD for advanced Hodgkin’s disease.
New Engl J Med 348(24):2386–2395
21. Borchmann P, Haverkamp H, Diehl V, Cerny T,
Markova J, Ho AD et al (2011) Eight cycles of
escalated-­
dose BEACOPP compared with four
cycles of escalated-dose BEACOPP followed by
four cycles of baseline-dose BEACOPP with or without radiotherapy in patients with advanced-stage
Hodgkin’s lymphoma: final analysis of the HD12 trial
of the German Hodgkin Study Group. J Clin Oncol
29(32):4234–4242
22. Engert A, Haverkamp H, Kobe C, Markova J, Renner
C, Ho A et al (2012) Reduced-intensity chemotherapy and PET-guided radiotherapy in patients with
advanced stage Hodgkin’s lymphoma (HD15 trial): a
randomised, open-label, phase 3 non-inferiority trial.
Lancet 379(9828):1791–1799
23. Hasenclever D, Loeffler M, Diehl V (1996) Rationale
for dose escalation of first line conventional chemotherapy in advanced Hodgkin’s disease. German
Hodgkin’s Lymphoma Study Group. Ann Oncol
7(Suppl 4):95–98
24. Tesch H, Diehl V, Lathan B, Hasenclever D, Sieber
M, Ruffer U et al (1998) Moderate dose escalation for advanced stage Hodgkin’s disease using
the bleomycin, etoposide, adriamycin, cyclophosphamide, vincristine, procarbazine, and prednisone
scheme and adjuvant radiotherapy: a study of the
German Hodgkin’s Lymphoma Study Group. Blood
92(12):4560–4567
25. Engert A, Diehl V, Franklin J, Lohri A, Dörken B,
Ludwig W-D et al (2009) Escalated-Dose BEACOPP
in the treatment of patients with advanced-stage
Hodgkin’s lymphoma: 10 years of follow-up of the
GHSG HD9 study. J Clin Oncol 27(27):4548–4554
26. von Tresckow B, Kreissl S, Goergen H, Brockelmann
PJ, Pabst T, Fridrik M et al (2018) Intensive treatment strategies in advanced-stage Hodgkin’s lymphoma (HD9 and HD12): analysis of long-term
13
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
Treatment of Advanced-Stage Hodgkin Lymphoma
survival in two randomised trials. Lancet Haematol
5(10):e462–ee73
Kobe C, Dietlein M, Franklin J, Markova J, Lohri A,
Amthauer H et al (2008) Positron emission tomography has a high negative predictive value for progression or early relapse for patients with residual
disease after first-line chemotherapy in advanced-­
stage Hodgkin lymphoma. Blood 112(10):3989–3994
Behringer K, Mueller H, Goergen H, Thielen I, Eibl
AD, Stumpf V et al (2013) Gonadal function and fertility in survivors after Hodgkin lymphoma treatment
within the German Hodgkin Study Group HD13 to
HD15 trials. J Clin Oncol 31(2):231–239
Federico M, Luminari S, Iannitto E, Polimeno G,
Marcheselli L, Montanini A et al (2009) ABVD compared with BEACOPP compared with CEC for the initial
treatment of patients with advanced Hodgkin’s lymphoma: results from the HD2000 Gruppo Italiano per lo
Studio dei Linfomi Trial. J Clin Oncol 27(5):805–811
Viviani S, Zinzani PL, Rambaldi A, Brusamolino
E, Levis A, Bonfante V et al (2011) ABVD versus
BEACOPP for Hodgkin’s lymphoma when high-dose
salvage is planned. New Engl J Med 365(3):203–212
Carde P, Karrasch M, Fortpied C, Brice P, Khaled H,
Casasnovas O et al (2016) Eight cycles of ABVD versus four cycles of BEACOPP escalated plus four cycles
of BEACOPP baseline in stage III to IV, International
prognostic score >/= 3, high-risk Hodgkin lymphoma:
first results of the phase III EORTC 20012 intergroup
trial. J Clin Oncol 34(17):2028–2036
Mounier N, Brice P, Bologna S, Briere J, Gaillard
I, Heczko M et al (2014) ABVD (8 cycles) versus
BEACOPP (4 escalated cycles >/= 4 baseline): final
results in stage III–IV low-risk Hodgkin lymphoma
(IPS 0-2) of the LYSA H34 randomized trial. Ann
Oncol 25(8):1622–1628
Merli F, Luminari S, Gobbi PG, Cascavilla N, Mammi
C, Ilariucci F et al (2016) Long-term results of the
HD2000 trial comparing ABVD versus BEACOPP
versus COPP-EBV-CAD in untreated patients with
advanced Hodgkin lymphoma: a study by Fondazione
Italiana Linfomi. J Clin Oncol 34(11):1175–1181
Skoetz N, Trelle S, Rancea M, Haverkamp H, Diehl
V, Engert A et al (2013) Effect of initial treatment
strategy on survival of patients with advanced-stage
Hodgkin’s lymphoma: a systematic review and network meta-analysis. Lancet Oncol 14(10):943–952
Skoetz N, Will A, Monsef I, Brillant C, Engert A,
von Tresckow B (2017) Comparison of first-line
chemotherapy including escalated BEACOPP versus chemotherapy including ABVD for people with
early unfavourable or advanced stage Hodgkin lymphoma. Cochr Datab Syst Rev 5:CD007941. https://
doi.org/10.1002/14651858
Hasenclever D, Diehl V (1998) A prognostic score for
advanced Hodgkin’s disease. International Prognostic
Factors Project on Advanced Hodgkin’s Disease. New
Engl J Med 339(21):1506–1514
Gallamini A, Hutchings M, Rigacci L, Specht L, Merli
F, Hansen M et al (2007) Early interim 2-[18F]fluoro-­
2-­
deoxy-D-glucose positron emission tomography
263
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
is prognostically superior to international prognostic score in advanced-stage Hodgkin’s lymphoma: a
report from a joint Italian-Danish study. J Clin Oncol
25(24):3746–3752
Meignan M, Gallamini A, Meignan M, Gallamini A,
Haioun C (2009) Report on the First International
Workshop on Interim-PET-Scan in Lymphoma.
Leukemia Lymphoma 50(8):1257–1260
Johnson P, Federico M, Kirkwood A, Fossa A, Berkahn
L, Carella A et al (2016) Adapted treatment guided
by interim PET-CT scan in advanced Hodgkin’s
Lymphoma. New Engl J Med 374(25):2419–2429
Gallamini A, Tarella C, Viviani S, Rossi A, Patti C,
Mule A et al (2018) Early chemotherapy intensification with escalated BEACOPP in patients with
advanced-stage Hodgkin lymphoma with a positive
interim positron emission tomography/computed
tomography scan after two ABVD cycles: long-term
results of the GITIL/FIL HD 0607 trial. J Clin Oncol
36(5):454–462
Borchmann P, Goergen H, Kobe C, Lohri A, Greil R,
Eichenauer DA et al (2018) PET-guided treatment in
patients with advanced-stage Hodgkin’s lymphoma
(HD18): final results of an open-label, international,
randomised phase 3 trial by the German Hodgkin
Study Group. Lancet 390(10114):2790–2802
Press OW, Li H, Schoder H, Straus DJ, Moskowitz
CH, LeBlanc M et al (2016) US intergroup trial of
response-adapted therapy for stage III to IV Hodgkin
lymphoma using early interim fluorodeoxyglucose-­
positron
emission
tomography
imaging:
Southwest Oncology Group S0816. J Clin Oncol
34(17):2020–2027
Zinzani PL, Broccoli A, Gioia DM, Castagnoli A,
Ciccone G, Evangelista A et al (2016) Interim positron emission tomography response-adapted therapy
in advanced-stage Hodgkin lymphoma: final results of
the phase II part of the HD0801 Study. J Clin Oncol
34(12):1376–1385
Kobe C, Goergen H, Baues C, Kuhnert G, Voltin CA,
Zijlstra J et al (2018) Outcome-based interpretation
of early interim PET in advanced-stage Hodgkin
lymphoma. Blood 132(21):2273–2279. https://doi.
org/10.1182/blood-2018-05-852129
Casasnovas RO, Brice P, Lazarovici J, Ghesquieres
H, Stamatoullas A, Dupuis J et al (2019) PET-adapted
treatment of patients with advanced Hodgkin lymphoma (AHL2011): final results of a randomised,
multi-centre, phase 3 study. Lancet 20(2):202–215
Biggi A, Gallamini A, Chauvie S, Hutchings M,
Kostakoglu L, Gregianin M et al (2013) International
validation study for interim PET in ABVD-treated,
advanced-stage Hodgkin lymphoma: interpretation
criteria and concordance rate among reviewers. J Nucl
Med 54(5):683–690
Borchmann P, Haverkamp H, Lohri A, Mey U, Kreissl
S, Greil R et al (2017) Progression-free survival of
early interim PET-positive patients with advanced
stage Hodgkin’s lymphoma treated with BEACOPP
escalated alone or in combination with rituximab
(HD18): an open-label, international, randomised
A. Fosså et al.
264
48.
49.
50.
51.
52.
53.
54.
phase 3 study by the German Hodgkin Study Group.
Lancet Oncol 18(4):454–463
Younes A, Gopal AK, Smith SE, Ansell SM,
Rosenblatt JD, Savage KJ et al (2012) Results of a
pivotal phase II study of brentuximab vedotin for
patients with relapsed or refractory Hodgkin’s lymphoma. J Clin Oncol 30(18):2183–2189
Younes A, Connors JM, Park SI, Fanale M, O’Meara
MM, Hunder NN et al (2013) Brentuximab vedotin combined with ABVD or AVD for patients with
newly diagnosed Hodgkin’s lymphoma: a phase
1, open-label, dose-escalation study. Lancet Oncol
14(13):1348–1356
Connors JM, Ansell SM, Fanale M, Park SI,
Younes A (2017) Five-year follow-up of brentuximab vedotin combined with ABVD or AVD for
advanced-stage classical Hodgkin lymphoma. Blood
130(11):1375–1377
Connors JM, Jurczak W, Straus DJ, Ansell SM, Kim
WS, Gallamini A et al (2018) Brentuximab vedotin
with chemotherapy for stage III or IV Hodgkin’s
Lymphoma. New Eng J Med 378(4):331–344
Chen RW, Gallamini A, Connors JM, Savage KJ,
Collins GP, Grigg A et al (2018) Brentuximab vedotin with chemotherapy for stage III or IV Hodgkin
­lymphoma (HL): Impact of cycle 2 PET result on
modified progression-free survival (mPFS). J Clin
Oncol 36:abstr 7539
Eichenauer DA, Plutschow A, Kreissl S, Sokler M,
Hellmuth JC, Meissner J et al (2017) Incorporation
of brentuximab vedotin into first-line treatment
of advanced classical Hodgkin’s lymphoma: final
analysis of a phase 2 randomised trial by the
German Hodgkin Study Group. Lancet Oncol
18(12):1680–1687
Roemer MG, Advani RH, Ligon AH, Natkunam
Y, Redd RA, Homer H et al (2016) PD-L1 and
55.
56.
57.
58.
59.
60.
PD-L2 genetic alterations define classical Hodgkin
lymphoma and predict outcome. J Clin Oncol.
34(23):2690–2697
Ramchandren R, Domingo-Domènech E, Rueda
A, Trněný M, Feldman TA, Lee HJ et al (2019)
Nivolumab for newly diagnosed advanced-stage
classic Hodgkin lymphoma: safety and efficacy in the phase II checkmate 205 study. J Clin
Oncol 37(23):1997–2007. https://doi.org/10.1200/
JCO.19.00315. PMID: 31112476
Loeffler M, Brosteanu O, Hasenclever D, Sextro
M, Assouline D, Bartolucci AA et al (1998) Meta-­
analysis of chemotherapy versus combined modality
treatment trials in Hodgkin’s disease. International
Database on Hodgkin’s Disease Overview Study
Group. J Clin Oncol 16(3):818–829
Andrieu JM, Yilmaz U, Colonna P, Loeffler M,
Brosteanu O, Hasenclever D (1999) MOPP versus
ABVD and low-dose versus high-dose irradiation
in Hodgkin’s disease at intermediate and advanced
stages: analysis of a meta-analysis by clinicians. J
Clin Oncol 17(2):730–732
Aleman BM, Raemaekers JM, Tirelli U, Bortolus R,
van’t Veer MB, Lybeert ML et al (2003) Involved-­
field radiotherapy for advanced Hodgkin’s lymphoma.
New Engl J Med 348(24):2396–2406
Ferme C, Mounier N, Casasnovas O, Brice P, Divine
M, Sonet A et al (2006) Long-term results and competing risk analysis of the H89 trial in patients with
advanced-stage Hodgkin lymphoma: a study by the
Groupe d'Etude des Lymphomes de l'Adulte (GELA).
Blood 107(12):4636–4642
Kriz J, Reinartz G, Dietlein M, Kobe C, Kuhnert G,
Haverkamp H et al (2015) Relapse analysis of irradiated patients within the HD15 trial of the German
Hodgkin Study Group. Int J Radiat Oncol 92(1):
46–53
Optimizing Decision Making
in Hodgkin Lymphoma
14
Susan K. Parsons, Joshua T. Cohen,
and Andrew M. Evens
Contents
14.1
Treatment Choices and Individualized Care
265
14.2
Treatment-Related Late Effects and Associated Human Cost
266
14.3
Risk, Impact, and Variability of Treatment-Related Late Effects
266
14.4
aucity of Harmonized Data to Guide Providers and Patients
P
Towards Individualized Treatment Choices
266
14.5
Disease Classification and Prognostication
267
14.6
14.6.1
14.6.2
Simulation Modeling
Low-Dose CT Scan for Lung Cancer
Diffuse Large B-Cell Lymphoma (DLBCL)
268
268
269
14.7
Decision Models in Hodgkin Lymphoma (HL)
270
14.8
Conclusion
271
References
272
14.1
S. K. Parsons (*)
Tufts Medical Center, Center for Health Solutions,
ICRHPS, Boston, MA, USA
e-mail: sparsons@tuftsmedicalcenter.org
J. T. Cohen
Tufts Medical Center, Center for the Evaluation of
Value and Risk in Health, ICRHPS, Boston,
MA, USA
e-mail: jcohen@tuftsmedicalcenter.org
A. M. Evens
Rutgers Robert Wood Johnson Medical School,
RWJBarnabas Health, New Brunswick, NJ, USA
Treatment Choices
and Individualized Care
Hodgkin lymphoma (HL) is one of the best curable cancers, particularly when presenting as
early-stage disease [1–3]. Although outcomes
differ by age, cure rates exceed 80–85% across
most stages and ages. Despite these overall excellent outcomes, there is no clear consensus regarding treatment recommendations across age
groups, and individual patients, with regard to
several treatment options, including which chemotherapy regimen to use, the optimal number of
chemotherapy cycles, and the role of sequential
adjunctive radiation therapy (RT) [1, 2, 4–13].
Furthermore, choices and debate over therapeutic
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_14
265
S. K. Parsons et al.
266
options have further expanded to include if and
how to integrate and use early/interim positron
emission tomography (PET) therapy to guide
treatment (i.e., response-adapted therapy), as
well as the optimum integration of novel therapeutics into frontline therapy [6, 14–17].
14.2
Treatment-Related Late
Effects and Associated
Human Cost
Critically, because the majority of newly diagnosed HL patients are young (median age 35
years) [18], curing disease can come at considerable “human cost,” including treatment-­related
toxicities and late effects (LE) (e.g., secondary
malignant neoplasms [SMN], cardiovascular disease [CVD]) and potential loss of young lives.
The incidence of CVD and SMN rises exponentially >20–30 years after treatment. Results from
analyses led by Dutch investigators and others
have shown that the risk of SMNs do not appear
to differ or be significantly lower over consecutive time periods [19–21]. Both Ng et al. [15] and
Castellino et al. [18] have highlighted increased
mortality among long-term HL survivors,
although both of these studies reflect the impact
of historical treatment approaches, including
extended field radiotherapy.
Additionally, cost-per-death analyses also
have shown that HL has the second highest cost
per death or lost productivity cost, behind only
malignant melanoma [22]. Further, productivity
analyses of cancer mortality have shown HL to
be the second most costly cancer in terms of lost
lifetime earnings [23]. In addition to economic
consequences, HL survivors also experience significantly compromised health-related quality of
life (HRQL) due to LEs [24].
14.3
isk, Impact, and Variability
R
of Treatment-Related Late
Effects
The risk of SMN depends on many clinical factors
(e.g., age at exposure, sex, stage) as well as several
treatment-related factors (e.g., chemotherapy [type
and number of cycles] and RT [dose and field]). A
recent Dutch analysis highlighted the impact of
radiation dose and field, sex, and smoking on the
risk of breast, lung, and other cancers [19]. Studies
have investigated the impact of age and sex on the
development of solid cancers after treatment for
HL [25] and the impact of sex and type of treatment (i.e., anthracycline chemotherapy ± radiation) on the incidence of major cardiac disease [26,
27]. However, it is not possible to use current population-level findings to reliably predict outcomes
of alternative therapies for specific, individual
patients and hence contribute to fully informed
decision-making.
14.4
aucity of Harmonized Data
P
to Guide Providers
and Patients Towards
Individualized Treatment
Choices
Helping clinicians assess and navigate alternative HL treatment options for individual patients
poses substantial challenges. First, ideal information is not available. Often, empirical data
for contemporary therapies is limited to relatively short-term follow-up with differences in
initial risk and response criteria driving therapy
and limited information about the risk and
severity of treatment-related LEs. While followup data from previous treatment eras offer
insights [28], treatment changes and improvements over time limit the relevance of existing
information. Second, the benefits and risks of
different therapies depend in part on individual
characteristics, such as patient age and sex,
among other factors. A recent HL position paper
by Travis and Ng et al. recommended development of c­ omprehensive risk prediction models
for LEs to customize treatment strategies [29].
There is a need to harmonize individual patient
data across age groups from recent trials and
existing datasets while establishing a data repository that facilitates incorporation of future data.
The past 15 years have seen publications of several clinical trials involving many pediatric and
adult HL patients with early-stage and advancedstage disease [3, 30]. However, each study exam-
14
Optimizing Decision Making in Hodgkin Lymphoma
ined a slightly different HL question and most
used different treatments. The result was a range
of distinct findings and hence a wide range of
therapeutic choices (Table 14.1).
Over the past 10 years, most HL studies worldwide have integrated PET-response-­
adapted
designs, an approach that directs the type and/or
amount of therapy based on PET scan results (positive vs. negative) early during the patient’s treatment course, usually after two chemotherapy
cycles [16, 33]. These PET-response-­adapted data
have significantly expanded the range of treatments
that providers and patients must consider in assessing treatment options for individual HL patients.
Taken together, there remains a multitude of
unanswered questions, especially for individual
HL patients, as exemplified by the index case
example above, including: (1) What is the efficacy of alternative treatments, and how do individual patient and disease characteristics
influence efficacy? (2) What is the HRQL impact
of each treatment option? (3) What are the incidence and severity of LEs (absolute risk for different SMNs and/or CVD), and how do these
outcomes depend on treatment, individual
patient characteristics, and disease characterisTable 14.1 HL case example
A 29-year-old female presents with increasing size
and number of lymph nodes and fatigue. Biopsy shows
nodular sclerosis HL. PET staging shows non-bulky
disease in right neck/supraclavicular, right hilum, and
bilateral axillary (i.e., stage IIA unfavorable). Past
family history includes father with myocardial
infarction at age 50.
Based on HL clinical study results, multiple valid
alternative treatment options exist for this common
case presentation, including:
– ABVD × 4–6 cycles (dependent on CT response)
without RT (NCI-C) [4, 5]
– ABVD × 3 with PET-based decision on RT
(i.e., none for CR) (RAPID) [6]
– ABVD × 4 cycles + RT based on PET (with
escalation to escalated BEACOPP for PET-2
positivity) (EORTC H10) [14, 15]
– 2 ABVD + 2 escalated BEACOPP + RT (GHSG
HD14) [15]
– ABVD × 2 with PET-2 and then ABVD × 2 (PET
negative) and BEACOPP × 2 (PET positive)
CALGB 50604 [31]
– 2 ABVD with PET-2 and then AVD × 6 (PET
negative) or BEACOPP × 6 (PET positive)
RATHL [32]
267
tics? (4) How does “real-world” HL data inform
treatment decisions in light of patient
preferences?
14.5
Disease Classification
and Prognostication
In adults, “early-stage” HL is often subdivided
into two categories designated “favorable” and
“unfavorable,” with the distinction made on the
basis of clinical factors and blood test results [3].
However, there are several different classifications
that have been developed and studied over the past
20 years in prospective HL clinical studies. The
criteria used by the German Hodgkin Study Group
(GHSG) and the European Organization for
Research and Treatment of Cancer (EORTC) differ with regard to several factors, such as number
of nodal groups. Furthermore, the “nodal maps”
differ, reflecting differences between GHSG and
EORTC clinical studies [14, 15, 34, 35]. In addition, clinical studies conducted in North America
have utilized different criteria to delineate earlystage disease, and some HL clinical trials have not
separated early-stage patients into different groups
[4, 5]. These staging definition differences can
influence the treatment patients receive and their
outcomes. Advanced disease has generally been
classified as Ann Arbor stage III and IV, but clinical trials have often included patients with high-­
risk stage II disease, such as those with
B-symptoms, involvement at multiple sites, and/or
bulky disease. The inclusion criteria have often
varied on a study-by-study basis, leading to substantial patient heterogeneity across HL studies.
Pediatric oncology research groups in the
United States and across the world have used different criteria than adult groups use to categorize
HL patients [10, 36]. While both pediatric and
adult groups rely on the Ann Arbor classification
system for staging, risk stratification has varied
within these risk groups. For example, some adult
studies classify patients with stage IIB disease
with bulk as having early-stage disease, while
pediatric trials currently designate patients as having advanced-stage disease based on inferior outcomes. Similarly, application of adult criteria
would classify pediatric patients with stage IIIA
disease as having advanced disease even though
S. K. Parsons et al.
268
these patients have had superior outcomes compared to other pediatric subgroups (e.g., stage IIIB,
IVA, IVB). Some pediatric trials do not include
these patients in advanced-stage studies, but,
rather, designate them to be at “intermediate” risk.
Prognosis in adult advanced-stage HL as
defined by the International Prognostic Index
(IPS) in 1998 includes measurements of albumin,
hemoglobin, sex, ages >45 years, stage IV, and the
presence of leukocytosis or lymphocytosis [37].
HL patients with higher IPS scores had inferior
treatment outcomes and were thus identified as
potentially requiring more intensive therapy. The
British Columbia Cancer Agency (BCCA) conducted an updated analysis of the IPS that showed
that the utility of the IPS was altered [38]. In this
analysis, the 5-year freedom from progression
(FFP) ranged from 62% to 88% and 5-year OS
ranged from 67% to 98% with a much narrower
range of outcomes for patients ages <65 years
(FFP ranging from 70% to 88% and 5-year OS
ranging from 73% to 98%). Furthermore, in a
multivariate regression analysis, which controlled
for all IPS factors, only age and hemoglobin level
retained independent significance.
Notably, no new and more comprehensive prognostic models have been developed for HL (early
stage or advanced stage) in more than 20 years.
Because age is an integral component of the original IPS, attempts have been made to develop and
validate a child-specific prognostic score, known as
CHIPS (Childhood Hodgkin International
Prognostic Score) [39]. The original testing found
that several factors were independent predictors of
event-free survival (EFS), including stage IV, large
mediastinal mass, low albumin, and fever. Further
validation in other cohorts of children and adolescents with advanced disease is underway.
14.6
Simulation Modeling
Statistical and simulation modeling offers a rigorous approach to systematically and explicitly
incorporate assumptions and information based
on multiple data sources to explore how alternative treatments affect outcomes of interest,
including LEs, survival, and quality-adjusted survival. Collectively, harmonization of independent
patient data from large, international prospective
studies and prominent cancer registries, along
with development of common data standards,
will establish robust “patient-specific” disease
progression and LE probabilities that may be harnessed for dynamic decision-making tools with
the expectation of ultimately improving outcomes across pediatric and adult HL.
Decision models have proved useful in connection with other diseases when treatment
options involve trade-offs, and the risks and benefits can vary substantially, depending on patient
characteristics. Here, we review the models
developed to evaluate measures to either help
prevent or treat two conditions: (1) lung cancer
and (2) diffuse large B-cell lymphoma (DLBCL).
14.6.1 L
ow-Dose CT Scan for Lung
Cancer
The National Lung Screening Trial (NLST) demonstrated that for patients at high risk for lung
cancer mortality, low-dose computed tomography (LDCT) reduces lung cancer mortality by 20
percent compared to screening by conventional
chest X-ray [40]. Cost-effectiveness analysis
revealed that compared to no screening, LDCT
accrues 0.02 quality adjusted life years (QALYs)
per person screened at an incremental cost of
$1631 [41]. The corresponding cost-effectiveness
ratio suggesting an outlay of $81,000 per QALY
gained represents “good value” relative to contemporary benchmarks for the United States [42].
Nonetheless, there remains the possibility that
more narrowly targeted selection of the population to be screened would accrue even greater
benefits and, hence, achieve a more favorable
cost-effectiveness ratio. Kovalchik et al. [43]
reported that after ranking the NLST population
by estimated lung cancer mortality risk, screening
prevented one lung cancer death for every 5276
individuals in the lowest-risk quintile, but that it
achieved the same benefit for every 161 screened
among the highest-risk quintile. The risk function
developed by Kovalchik et al. therefore offers an
approach for reducing the amount of screening
needed to achieve the same mortality reduction.
Kumar et al. [44] investigated the efficiency of
targeting individuals at even higher risk for lung
cancer mortality than the NLST population as a
14
Optimizing Decision Making in Hodgkin Lymphoma
whole. Using a different risk model than Kovalchik
et al., Kumar et al. reported similar efficiency
gains for reducing lung cancer mortality.
Screening top decile individuals yielded a nearly
eightfold gain in averted deaths per person
screened, compared to screening of individuals in
the bottom decile. Assessment using other outcome measures yielded less impressive efficiency
gains. For life years gained, the benefit per person
screened was 3.6 times greater for top mortality
risk decile individuals, compared to the bottom
decile. For quality-adjusted life year gains
(QALYs), the corresponding ratio was 2.4.
Finally, the cost-effectiveness of screening
improved across the risk deciles by an even
smaller relative margin, from $75,000 per QALY
gained in the lowest-risk decile to $53,000 per
QALY gained in the highest-risk decile, a ratio of
approximately 1.4.
The broad range of efficiency gains for different outcome measures illustrates both the
strengths and limitations of risk targeting. When
the targeting criterion—lung cancer mortality
risk, in this case—matches the outcome measure,
targeting vastly improves efficiency. On the other
hand, when the outcome measure is less tightly
related to the targeting measure, potential efficiency gains can decrease. Kumar et al. note, for
example, that in the NLST cohort, higher-risk
individuals were older, had greater smoking exposure, and were more likely to have a preexisting
diagnosis of chronic obstructive pulmonary disease [44]. Because the characteristics making
these individuals “high risk” also reduce life
expectancy, targeting is less effective at maximizing life year gains. Likewise, mortality risk is
inversely associated with future quality of life and
positively associated with higher future care costs.
Those factors further mitigate the efficiency gains
from mortality risk targeting measured in terms of
QALY gains and cost-effectiveness.
14.6.2 D
iffuse Large B-Cell
Lymphoma (DLBCL)
Because there are multiple treatments for patients
with DLBCL, and because treatments vary in
terms of their intensity and side effects, an accurate prognosis is crucial to identifying a course of
269
care that appropriately accounts for a patient’s
risks and benefits.
In recent decades, clinicians have relied on the
International Prognostic Index (IPI) to characterize risk [45]. The IPI produces a risk score ranging from 0 to 5 based on a series of dichotomized
risk factors, including age (less than 60 vs. 60 or
older), number of extranodal sites (0–1 vs. 2 or
more), Ann Arbor stage (I or II vs. III or IV), lactate dehydrogenase levels (not elevated vs. elevated), and Eastern Cooperative Oncology Group
performance status (2 or less vs. greater than 2).
Incorporation of additional prognostic characteristics has improved the prognostic accuracy of
the IPI, but limitations remain, including the
dichotomous characterization of inputs and the
tool’s semi-qualitative characterization of risk
that does not specify probabilities for key outcomes such as mortality.
To address these limitations, Biccler et al. [45]
developed a model to predict overall survival and
event-free survival as a function of both categorical characteristics (e.g., sex, Ann Arbor stage,
presence or absence of B symptoms, among others) and continuous values (e.g., log leukocyte
count, hemoglobin level, among others). The prediction reflects a weighted average of statistical
models, with weights selected to maximize prediction accuracy.
The authors have made the model available on
the Internet (https://lymphomapredictor.org/).
The results show overall survival (compared to
background survival) and event-free survival,
both of which are projected over a period of five
years. Because this model incorporates finegrained individual characteristics and reports
outcome probabilities over time, it represents a
substantial improvement over earlier prognostic
tools. Nonetheless, it has two key limitations.
First, its projections are limited to a period of five
years. That limitation reflects the extent of the
follow-up in the population data used to build the
model. Second, the model does not describe how
alternative treatments influence outcomes. While
clinicians and patients may infer that higher risks
warrant more intensive treatment, the model does
not quantify the resulting trade-offs.
Altogether, trade-offs, in the form of adverse
events and resource costs, are common. Typically,
these downside impacts do not depend on the size
S. K. Parsons et al.
270
of the potential benefit. The size of the potential
benefit, and hence the magnitude of the net benefit, often depends on how big the baseline risk is.
As a result, targeting individuals at highest risk
for disease or severe outcomes often increases
efficiency. The effectiveness of this strategy
depends on the strength of the association
between the risk stratification measure and the
benefit measure. Prognostic risk models can help
clinicians and patients weigh treatment alternatives, but these models can be limited by their
time horizon and the extent to which they incorporate the impact of alternative therapies.
14.7
ecision Models in Hodgkin
D
Lymphoma (HL)
Given the varying treatment approaches and
their trade-offs relative to disease control and
LE risks and the impact of individual patient
characteristics, such as gender and age at the
time of exposure, there has been considerable
interest in the development of decision models
for newly diagnosed HL. Our initial model of
early-stage HL utilized published, group-level
data from recent clinical trials to estimate simulated short-term and long-term outcomes [46].
We began with the development of a detailed
disease map (Fig. 14.1), which highlights the
health states through which a patient can move
once diagnosed with HL. Based on best available information, we estimated the probability
of transitioning from one health state to the next
and the HRQL of each health state in the form
of utility weights.
To test the model, we created two hypothetical
cases that differed with regard to gender, disease
location, and extent of disease. We then compared the projected outcomes (life expectancy
and QALYs) for each patient for each of two
treatment modalities—chemotherapy alone and
combined modality therapy [46]. Sensitivity
analyses explored the impact on projected c­ linical
outcomes of age at diagnosis and the assumed
incidence and severity of late effects. The purpose of this initial model was not to identify
which modality might be definitively superior to
the other, but, rather, to illustrate that treatment
recommendations should reflect patient factors,
4
B
5
Relapse
0.71
A
Diagnosis &
Treatment
9
C1
1
Cured With
Relapse
0.74
C3
6
C2
Cured
Without
Relapse
0.80
At Risk
0.77
2
7
D
Cured
With Late
Effects
0.67-0.73
Dead
0.0
10
8
3
Fig. 14.1 Utility Weight Simulation Model: The State
Transition Diagram. The bubbles represent individual
health states. The value within each bubble is the utility
weight (or health-related quality of life impact) of that
health state. The arrows represent transition pathways
between states. Scaling each year of survival by that
year’s utility weight (specified for each health state in the
figure) and then summing the quality-adjusted years
yields quality-adjusted survival
14
Optimizing Decision Making in Hodgkin Lymphoma
Fig. 14.2 The State
Transition Diagram—
therapeutic-response
enhanced model.
Individuals in State A
can proceed to States B
or C2. The transition
rates depend on whether
the patient experiences a
rapid or slow response
to initial therapy—i.e.,
whether the individual
progresses from substate
A-i to A-ii, or to A-iii
271
B
A-ii
At-risk
State “A”
Relapse
0.71
1
RER
C2
A-i
2
Preresponse
data
A-iii
Cured
Without
Relapse
0.80
SER
disease characteristics, and the outcome preferences of the patient and his/her provider.
Decision models such as these can also be
adapted as new information becomes available.
For example, consider the use of early PET-based
response. In the figure below, we have added the
PET-adapted response as a new health state (that
is, rapid early response or slow early response)
(Fig. 14.2). Because addition of this new health
state requires revision of the probabilities downstream, such as risk of relapse, the revised decision model can be run to estimate updated clinical
outcomes.
To extend this one step further, one could utilize this type of model as patients transition from
one phase of care to another, namely, from active
treatment to active surveillance or active surveillance to survivorship. By incorporating emerging
information, patients and their providers can
refine ongoing care needs and clarify areas of
likely risk and uncertainty.
The development of these types of dynamic
decision models requires individual patient data
from large numbers of patients to account for differences across patients in terms of their demographic characteristics and disease factors.
Moreover, model development depends on identifying data projecting the impact of contemporary treatment on short- and long-term outcomes,
including toxicity, death, relapse, and LEs. Data
must also be updated, as additional information
becomes available (e.g., from new trials, or from
further follow-up of existing trials) and as new
therapies are introduced.
Critical to the implementation and dissemination of these tools is an understanding of how
patients, caregivers, and providers would use
such models in the real world, what concerns
they have, what kind of decision support they
need, and what outcomes they are interested in.
As noted, our first version of the model estimated life expectancy and QALYs, but these outcomes can be modified to reflect what is salient
and/or accessible to different stakeholders.
14.8
Conclusion
Given the success of frontline treatments and the
ability to salvage the majority of HL patients
after disease progression or recurrence, the overall survival of HL is high. However, this survival
comes at a cost to patients in the form of LEs,
which can alter both the length and quality of survivorship. To reduce downstream LE risk, modifications have been made in frontline therapy,
including: changes in indications for radiation,
reduction in radiation dose and field among those
S. K. Parsons et al.
272
receiving treatment, risk stratification to determine need for either dose reduction or dose escalation to optimize outcomes, and incorporation of
novel agents, initially in the salvage setting and
more recently in frontline therapy.
Through data sharing and international collaboration, including across pediatric and adult
specialties, we can create robust and nimble decision models to guide our patients and their families, alongside their providers, to enhance and
optimize the difficult decisions that affect acute
and long-term outcomes.
12.
13.
14.
References
1. Giulino-Roth L et al (2015) Current approaches in the
management of low risk Hodgkin lymphoma in children and adolescents. Br J Haematol 169(5):647–660
2. Armitage JO (2010) Early-stage Hodgkin’s lymphoma. N Engl J Med 363(7):653–662
3. Evens AM, Hutchings M, Diehl V (2008) Treatment
of Hodgkin lymphoma: the past, present, and future.
Nat Clin Pract Oncol 5(9):543–556
4. Meyer RM et al (2005) Randomized comparison of ABVD chemotherapy with a strategy that
includes radiation therapy in patients with limited-stage Hodgkin’s lymphoma: National Cancer
Institute of Canada Clinical Trials Group and the
Eastern Cooperative Oncology Group. J Clin Oncol
23(21):4634–4642
5. Meyer RM et al (2012) ABVD alone versus radiationbased therapy in limited-stage Hodgkin’s lymphoma.
N Engl J Med 366(5):399–408
6. Radford J et al (2015) Results of a trial of PET-directed
therapy for early-stage Hodgkin’s lymphoma. N Engl
J Med 372(17):1598–1607
7. Percival ME, Hoppe RT, Advani RH (2014) Bulky
mediastinal classical Hodgkin lymphoma in young
women. Oncology (Williston Park) 28(3):253-6–25860. C3
8. Crump M et al (2015) Evidence-based focused
review of the role of radiation therapy in the treatment of early-stage Hodgkin lymphoma. Blood
125(11):1708–1716
9. Hay AE et al (2013) An individual patient-data comparison of combined modality therapy and ABVD
alone for patients with limited-stage Hodgkin lymphoma. Ann Oncol 24(12):3065–3069
10. Wolden SL et al (2012) Long-term results of CCG
5942: a randomized comparison of chemotherapy
with and without radiotherapy for children with
Hodgkin’s lymphoma—a report from the Children’s
Oncology Group. J Clin Oncol 30(26):3174–3180
11. Nachman JB et al (2002) Randomized comparison of
low-dose involved-field radiotherapy and no radio-
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
therapy for children with Hodgkin’s disease who
achieve a complete response to chemotherapy. J Clin
Oncol 20(18):3765–3771
Straus DJ et al (2004) Results of a prospective randomized clinical trial of doxorubicin, bleomycin,
vinblastine, and dacarbazine (ABVD) followed by
radiation therapy (RT) versus ABVD alone for stages
I, II, and IIIA nonbulky Hodgkin disease. Blood
104(12):3483–3489
Laskar S et al (2004) Consolidation radiation after
complete remission in Hodgkin’s disease following
six cycles of doxorubicin, bleomycin, vinblastine,
and dacarbazine chemotherapy: is there a need? J Clin
Oncol 22(1):62–68
Raemaekers JM et al (2014) Omitting radiotherapy in
early positron emission tomography-negative stage I/
II Hodgkin lymphoma is associated with an increased
risk of early relapse: Clinical results of the preplanned
interim analysis of the randomized EORTC/LYSA/
FIL H10 trial. J Clin Oncol 32(12):1188–1194
Andre MP et al (2017) Early positron emission
tomography response-adapted treatment in stage
I and II Hodgkin lymphoma: final results of the
randomized EORTC/LYSA/FIL H10 trial. J Clin
Oncol 35(16):1786–1794. https://doi.org/10.1200/
JCO.2016.68.6394
Evens AM, Kostakoglu L (2014) The role of FDGPET in defining prognosis of Hodgkin lymphoma for
early-stage disease. Blood 124(23):3356–3364
Olszewski AJ, Shrestha R, Castillo JJ (2015)
Treatment selection and outcomes in early-stage classical Hodgkin lymphoma: analysis of the National
Cancer Data Base. J Clin Oncol 33(6):625–633
Available from https://seer.cancer.gov/csr/1975_2015/.
Schaapveld M et al (2015) Second cancer risk up to
40 after treatment for Hodgkin’s lymphoma. N Engl J
Med 373(26):2499–2511
Aleman BM et al (2007) Late cardiotoxicity
after treatment for Hodgkin lymphoma. Blood
109(5):1878–1886
van Leeuwen FE, Ng AK (2016) Long-term risk of
second malignancy and cardiovascular disease after
Hodgkin lymphoma treatment. Hematology Am Soc
Hematol Educ Program 2016(1):323–330
Hanly P, Soerjomataram I, Sharp L (2015) Measuring
the societal burden of cancer: the cost of lost productivity due to premature cancer-related mortality in
Europe. Int J Cancer 136(4):E136–E145
Bradley CJ et al (2008) Productivity costs of cancer
mortality in the United States: 2000–2020. J Natl
Cancer Inst 100(24):1763–1770
Linendoll N et al (2016) Health-related quality of life
in Hodgkin lymphoma: a systematic review. Health
Qual Life Outcomes 14(1):114
Hodgson DC et al (2007) Long-term solid cancer risk
among 5-year survivors of Hodgkin’s lymphoma. J
Clin Oncol 25(12):1489–1497
Myrehaug S et al (2008) Cardiac morbidity following modern treatment for Hodgkin lymphoma: supra-
14
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
Optimizing Decision Making in Hodgkin Lymphoma
additive cardiotoxicity of doxorubicin and radiation
therapy. Leuk Lymphoma 49(8):1486–1493
Myrehaug S et al (2010) A population-based study
of cardiac morbidity among Hodgkin lymphoma
patients with preexisting heart disease. Blood
116(13):2237–2240
Castellino SM et al (2011) Morbidity and mortality in
long-term survivors of Hodgkin lymphoma: a report
from the Childhood Cancer Survivor Study. Blood
117(6):1806–1816
Travis LB et al (2012) Second malignant neoplasms
and cardiovascular disease following radiotherapy. J
Natl Cancer Inst 104(5):357–370
Diefenbach CS et al (2017) Hodgkin lymphoma: current status and clinical trial recommendations. J Natl
Cancer Inst 109:4
Straus DJ et al (2018) CALGB 50604: risk-adapted
treatment of nonbulky early-stage Hodgkin lymphoma
based on interim PET. Blood 132(10):1013–1021
Johnson P et al (2016) Adapted treatment guided by
interim PET-CT scan in advanced Hodgkin’s lymphoma. N Engl J Med 374(25):2419–2429
Coyle M, Kostakoglu L, Evens AM (2016) The evolving role of response-adapted PET imaging in Hodgkin
lymphoma. Ther Adv Hematol 7(2):108–125
von Tresckow B et al (2012) Dose-intensification in
early unfavorable Hodgkin’s lymphoma: final analysis of the German Hodgkin Study Group HD14 trial. J
Clin Oncol 30(9):907–913
Engert A et al (2010) Reduced treatment intensity in
patients with early-stage Hodgkin’s lymphoma. N
Engl J Med 363(7):640–652
Keller FG et al (2018) Results of the AHOD0431
trial of response adapted therapy and a salvage strategy for limited stage, classical Hodgkin lymphoma: a
report from the Children’s Oncology Group. Cancer
124(15):3210–3219
273
37. Hasenclever D, Diehl V (1998) A prognostic score for
advanced Hodgkin’s disease. International Prognostic
Factors Project on Advanced Hodgkin’s Disease. N
Engl J Med 339(21):1506–1514
38. Moccia AA et al (2012) International Prognostic Score
in advanced-stage Hodgkin’s lymphoma: altered utility in the modern era. J Clin Oncol 30(27):3383–3388
39. Schwartz CL et al (2017) Childhood Hodgkin
International Prognostic Score (CHIPS) Predicts
event-free survival in Hodgkin lymphoma: a report
from the Children’s Oncology Group. Pediatr Blood
Cancer 64:4
40. Aberle DR et al (2011) Reduced lung-cancer mortality with low-dose computed tomographic screening.
N Engl J Med 365(5):395–409
41. Black WC et al (2014) Cost-effectiveness of CT
screening in the National lung screening trial. N Engl
J Med 371(19):1793–1802
42. Neumann PJ, Cohen JT, Weinstein MC (2014)
Updating cost-effectiveness—the curious resilience
of the $50,000-per-QALY threshold. N Engl J Med
371(9):796–797
43. Kovalchik SA et al (2013) Targeting of low-dose CT
screening according to the risk of lung-cancer death.
N Engl J Med 369(3):245–254
44. Kumar V et al (2018) Risk-targeted lung cancer
screening: a cost-effectiveness analysis. Ann Intern
Med 168(3):161–169
45. Biccler J et al (2018) Simplicity at the cost of predictive accuracy in diffuse large B-cell lymphoma:
a critical assessment of the R-IPI, IPI, and NCCNIPI. Cancer Med 7(1):114–122
46. Parsons SK et al (2018) Early-stage Hodgkin lymphoma in the modern era: simulation modelling to
delineate long-term patient outcomes. Br J Haematol
182(2):212–221
Part III
Special Clinical Situations
Pediatric Hodgkin Lymphoma
15
Georgina W. Hall, Cindy Schwartz, Stephen Daw,
and Louis S. Constine
Contents
15.1
15.1.1
15.1.2
15.1.2.1
15.1.2.2
15.1.2.3
15.1.2.4
15.1.3
15.1.4
15.1.4.1
15.1.4.2
15.1.4.3
15.1.4.4
15.1.4.5
15.1.4.6
15.1.4.7
15.1.5
15.1.5.1
15.1.5.2
15.1.5.3
15.1.5.4
15.1.6
Introduction Comparison of Pediatric/Adolescent Vs. Adult HL Classical Pediatric Hodgkin Lymphoma (PHL) Overall Strategies Low-Risk (Early Favorable) Disease Intermediate- and High-Risk (Advanced, Unfavorable) Disease Future Considerations in Classical Pediatric and Adolescent HL Nodular Lymphocyte-­Predominant HL (NLPHL) Recurrence, Relapse, and Salvage in PHL Introduction Standard-Dose Salvage Chemotherapy Regimens Prognostic Factors at Relapse in Pediatric HL: Standard-­Dose
Chemoradiotherapy Vs. High-Dose Chemotherapy/Stem Cell
Transplantation Role of Radiotherapy in Relapsed Hodgkin Lymphoma High-Dose Chemotherapy and Autologous Stem Cell Transplant High-Dose Chemotherapy and Allogeneic Stem Cell Transplantation Brentuximab Vedotin and Checkpoint Inhibitors Late Effects Cardiac Toxicities Pulmonary Toxicities Thyroid Toxicities Secondary Malignancies Summary/Future Directions References G. W. Hall (*)
Pediatric and Adolescent Haematology/Oncology
Unit, Children’s Hospital, John Radcliffe Hospital,
Oxford University Hospitals NHS Trust, Oxford, UK
e-mail: georgina.hall@paediatrics.ox.ac.uk
C. Schwartz
Division of Pediatric Hematology Oncology and
BMT, Medical College of WI, Milwaukee, WI, USA
e-mail: cschwartz@mcw.edu
278
278
278
278
280
282
285
285
286
286
287
287
288
288
289
289
290
290
290
290
291
291
292
S. Daw
Children and Young People’s Cancer Services,
Division of Paediatrics, University College Hospital,
London, UK
e-mail: stephendaw@nhs.net
L. S. Constine
Department of Radiation Oncology and Pediatrics,
University of Rochester Medical Center,
Rochester, NY, USA
e-mail: louis_constine@urmc.rochester.edu
© Springer Nature Switzerland AG 2020
A. Engert, A. Younes (eds.), Hodgkin Lymphoma, Hematologic Malignancies,
https://doi.org/10.1007/978-3-030-32482-7_15
277
G. W. Hall et al.
278
15.1
Introduction
15.1.1 Comparison of Pediatric/
Adolescent Vs. Adult HL
Pediatric/young adult Hodgkin lymphoma (HL)
is one of the few childhood malignancies that
shares aspects of its biology and natural history
with an adult cancer. Historically, children were
thought to have a worse prognosis than adults due
to antiquated treatment approaches that were initially designed to mitigate toxicities in children.
It is now clear that effective therapy provides
similar or even superior outcomes in children/
young adults. A comparison of the demographics
of clinical presentations of pediatric/adolescent
HL compared with adult HL is presented in
Table 15.1. The first of the bimodal incidence
peaks in HL occurs in teenagers and young adults
(15–25-year age group). HL represents less than
5% of malignancies in children under the age of
15 years. In contrast, it represents 16–20% of
malignancies in adolescents, making it the most
common malignancy of this age group.
Childhood HL is biologically indistinguishable from HL of young and middle-aged adults
other than the relative incidence of specific disease histologies (Table 15.1). Mixed cellularity
(MC) and nodular lymphocyte predominant
(nLP) HL are the common types of HL in the preadolescent child; adolescents and young adults
are most frequently (85%) afflicted with nodular
sclerosing (NS) HL [3]. Only a third of children
will have advanced disease; approximately 25%
will have B symptoms. The incidence of HL with
adverse features increases with age. Although
there were no discernable differences in clinical
presentation, response to therapy, or long-term
outcome for adolescents (16–21 years) vs. young
adults (22–45 years) treated similarly for HL [4],
the treatment of children/adolescents and adults
has diverged over the years primarily due to concerns about the late adverse effects of therapy.
15.1.2 C
lassical Pediatric Hodgkin
Lymphoma (PHL)
15.1.2.1 Overall Strategies
The adverse consequences of therapy have
driven the pediatric treatment paradigm of care.
Clinical trials for pediatric and adolescent HL
have been designed to both reduce long-term
organ injury and increase efficacy. Pediatric
oncologists responded first to developmental
issues in the young child and later to the long-
Table 15.1 Demographic and clinical characteristics at presentation of pediatric HL (modified from Refs. [1, 2])
Age range (years)
Prevalence of HL cases (%)
Gender
Male/female
Histology
Nodular sclerosis (%)
Mixed cellularity (%)
Lymphocyte depleted (%)
NLPHL (%)
EBV associated
Other risk factors
Childhood HL
≤14
10–12
2–3:1
AYA HL
15–35
50
1:1–1.3:1
40–45
30–45
0–3
8–20
65–80
10–25
1–5
2–8
27–54%
Risk factors: male, younger age,
mixed cellularity histology,
economically disadvantaged countries
Lower SES increasing family size
20–25%
Stage at presentation
30–35% with stage III or IV disease,
25% with B symptoms
Relative survival rates at
5 years
94% (<20 years)
Higher SES, smaller family
size, early birth order
40% with stage III or IV
disease, 30–40% with B
symptoms
90% (<50 years)
AYA adolescents and young adults, IPS International Prognostic Score, SES socioeconomic status
Adult HL
≥35
34–40%
15
Pediatric Hodgkin Lymphoma
279
term treatment consequences in all young survivors in the design of treatment approaches.
Recognition of musculoskeletal hypoplasia in
young children with HL treated with high-dose
radiation such as shortened sitting height, thin
necks, and narrow shoulders and chest [5–8] precipitated the development of pediatric-specific
regimens for HL. Combined-­
modality treatments, even for low-stage disease, allowed for
the reduction of radiation dose [9] and field size,
thus sparing normal structures (Fig. 15.1). This
strategy for care was extended to older children
and adolescents when ­hypothyroidism [11, 12],
secondary cancers, and valvular and atherosclerotic heart disease [13, 14] were also found to be
attributable to high-dose radiation.
Low-dose radiation of 15–25 Gy has been the
standard in childhood and adolescent HL for
a
b
c
Proportional Reduction in Mean Dose
1.2
Normalized mean dose
decades. This reduced the potential for longterm risk without adversely impacting event-free
survival. A convergence of treatment approaches
for adults and children/adolescents may be
emerging as recent adult trials have begun to
address these issues and reduce radiation doses.
With overall survival over 90%, the quality of
survival becomes paramount.
Early response to therapy was recognized [15,
16] as highly predictive of outcome. In Europe
and the United States, response-based, risk-­
adapted approach to treating HL [17] allows therapy to be tailored to each individual, within the
context of clinical trials. Dose-dense regimens
[17] used are similar to those used by adult
groups [18, 19], but the pediatric algorithms use
the enhanced efficacy to support reduction of
therapy.
1
0.8
Mantle 36Gy
0.6
IFRT 21Gy
INRT 21Gy
0.4
0.2
0
Bilateral Breast
Bilateral Lungs
Heart
Thyroid
Organ Exposed
Fig. 15.1 CT-based planning images depicting a historic
mantle RT, compared to standard involved field radiation
treatment (IFRT) and involved-node RT (INRT) for a
patient with stage I disease involving the mediastinum.
The postchemotherapy volume of initially involved
paratracheal nodes is depicted in dark red and the cardiac
silhouette is also evident. (a–c) Demonstration of the
reduction in dose to breast, lung, heart, and thyroid for the
female patient shown in (a) from Mantle 36Gy to IFRT
21Gy to INRT 21Gy. From Hodgson et al. [10]
G. W. Hall et al.
280
Table 15.2 Risk groups employed by selected pediatric study groups [20]
Study group
Children’s Oncology
Group [21, 22]
EuroNet-PHL-C1,
C2 [23]
St. Jude/Stanford/
Dana-Farber
Risk features (RF)
Categorized as favorable or
unfavorable risk by IPS
Low risk
IA/IIA no
bulk/no LMA
IA/Ba
IIAa
IA/IIA no
bulk
Intermediate/early
unfavorable risk
All others IIB, IIIA
IVA
IIA,
IIB (no E), IIIA (no E)
IA/IIA (RF), I
IB
IIIA
IIII
High risk
IIIB, IVB
IIEB IIIEA/
IIIB IV
IIB, IIIB,
IV.
No bulk, ESR < 30 mm/h
a
15.1.2.2
ow-Risk (Early Favorable)
L
Disease
Although there have been differing definitions of
low-risk disease (Table 15.2), risk-adapted
approaches aim to define a cohort of patients that
is curable with minimal therapy. Treatment group
allocation, risk stratification, and response assessment vary according to each study group
(Table 15.2), but all treatment groups define low
risk based on stage and bulky disease. Children
and adolescents with NLPHL are increasingly
being treated with surgery alone or using low-­
dose regimens separate from those used for the
treatment of classical HL.
In the decade following the introduction of
MOPP, secondary leukemia and sterility emerged
as significant concerns [24–27]. During the
1980s, alkylator exposure and leukemia risk were
reduced by alternating MOPP and ABVD [28,
29]. The goal was to avoid reaching thresholds of
toxicity for any specific agent. The Pediatric
Oncology Group (POG) compared four cycles of
MOPP/ABVD plus 25.5 Gy to six cycles of chemotherapy alone without detecting differences in
efficacy [15]. However, the profound sensitivity
of the testes to procarbazine continued to cause
sterility in boys, even with only two cycles of
procarbazine-containing chemotherapy [30].
Although early attempts to avoid procarbazine
were unsuccessful [31], more recent regimens
have achieved this goal [17].
ABVD is used routinely in adults [32], but also
has not been standard of care in children.
Successful regimens have been devised by the
German Paediatric Oncology Hodgkin’s Group
(GPOH) [33] using OEPA (vincristine, etopo-
side, prednisone, and doxorubicin) in males
(Table 15.3), by the French Society of Pediatric
Oncology [36] using EBVP (etoposide, bleomycin, vincristine, prednisone), by Donaldson et al.
[42] using VAMP (vincristine, doxorubicin, methotrexate, and prednisone), and by the Children’s
Oncology Group (COG) using ABVE (doxorubicin, bleomycin, vincristine, etoposide) [43] and
ABV-PC [41] all avoiding the use of procarbazine.
With these approaches, EFS of 88–92% can be
achieved without significant radiation or alkylator
toxicity. Patients treated on these newer regimens
receive less than 200 mg/m2 of doxorubicin plus or
minus 20–25 Gy of involved-field radiation.
The traditional approach of most pediatric HL
treatment groups has been to use combined-­
modality therapy. Currently, these study groups
are involved in evaluating methods to define low-­
risk patients who may be cured without radiotherapy, i.e., with chemotherapy alone. However,
patients with early-stage HL treated with chemotherapy alone most frequently relapse in the initially involved lymph node(s) [44]. Therefore, an
effort has also been made to reduce the radiation
field size by including only the initially involved
lymph node(s)—so-called involved node-­
radiation (INRT) [45]. The complexity of defining
the field for INRT has led to the development of
an alternative approach termed “involved-site
radiation therapy” (ISRT) [46–48]. This is a
modification of IFRT, recommended for patients
who when optimal pre-chemotherapy imaging
(PET-CT in a position similar to what will be used
at the time of radiation therapy) is not available
that would be necessary for INRT treatment planning. Because the delineation of the area of
15
Pediatric Hodgkin Lymphoma
281
Table 15.3 Treatment results for early, favorable pediatric HL
Group or
institution
Combined-­
modality trials
US CCG 5942
Patients
(n)
Stage
Chemotherapy
Radiation
(Gy). field
Survival (%)
DFS,
EFS, Follow-up
or
interval
Overall RFS (years)
294
IA/B, IIA
4 COPP/ABV
21, IF
100
97
SFOP MDH-90
171
I − II
20, IF
97.5
91
27
I − II
78
5
GPOH-HD95
326,224
I, IIA IIB,
IIIA
4 VBVP, good
responders
4 VBVP 1–2
OPPA, poor
responders
2 OEPA or 2
OPPA
above + COPP
3
10
5
GPOH-HD2002
195,139
IA, 1B,
IIA, IE,
IIB,IIAE,
IIIA
2 OEPA or 2
OPPA
Above +2
COPP or 2
COPDAC
Chemotherapy
alone
US CCG 5942
106
CS IA/B,
IIA
IA, IIA
IA, IIA
Response based
RT
AHOD0431
[D,E]
St. Jude
consortium[C]
278
88
20, IF
References
[34, 35]
[36]
CR: No RT
PR: 20 IF
(10–15 Gy
boost)
20 ± 10–15
IF
99 97
94
88
5
[37, 38]
99.5
98.5
92
88
5
[39, 40]
4 COPP/ABV
CR: None
100
91
3
[34]
4 AV-PC
4 VAMP
21 IF if PR
25.5 IF/
none if
Early CR
99
100
80
89
4
5
[41]
ABVD Adriamycin, bleomycin, vinblastine, and dacarbazine, AEIOP Italian Association of Hematology and Pediatric
Oncology, CCG Children’s Cancer Group, ChlVPP chlorambucil, vinblastine, procarbazine, and prednisolone, COPP
cyclophosphamide, vincristine (Oncovin), prednisone, and procarbazine, COPP/ABV cyclophosphamide, vincristine
(Oncovin), procarbazine, prednisone, Adriamycin, bleomycin, and vinblastine, CR complete response, CS clinical
stage, EF extended field, EFS event-free survival, HD Hodgkin’s disease, IF involved field, MDH multicenter trial, MH
multicenter Hodgkin’s trial, MOPP nitrogen mustard, vincristine (Oncovin), procarbazine, and prednisone, M/T mediastinal/thoracic ratio, OEPA vincristine (Oncovin), etoposide, prednisone, and Adriamycin, OPA vincristine (Oncovin),
prednisone, and Adriamycin, OPPA vincristine (Oncovin), procarbazine, prednisolone, and Adriamycin, PR partial
response, PS pathologic stage, R regional, RFS relapse-free survival, RT radiotherapy, SFOP French Society of Pediatric
Oncology, VAMP vinblastine, Adriamycin, methotrexate, and prednisone, VBVP vinblastine, bleomycin, etoposide
(VP-16), and prednisone, AVPC doxorubicin, vincristine, prednisone, etoposide
Mediastinal thoracic ratio < 0.33, lymph node <6 cm
involvement is less precise, a somewhat larger
treatment volume is treated than with INRT, but
less than traditionally used with IFRT. Other radiation techniques that are contemporary and reduce
the treatment volume include intensity-­modulated
radiation therapy, deep inspiration breath holding
(to reduce the volumes of the lung and heart that
might be exposed), and protons [49].
Nachman et al. showed an increased relapse
rate in patients who did not receive radiation
despite achieving CR at the end of chemotherapy [34, 35]. Late-response evaluation may not
have identified the optimal cohort for reduction
of radiation. Early response may better define
the profoundly chemotherapy-sensitive patient
who does not need radiation. Based on the
282
excellent outcomes of low-risk HL patients
achieving CR after two cycles of chemotherapy
[15], recent trials in the COG, the St. Jude/
DFCI/Stanford Consortium, and the EuroNet
PHL group [50, 51], have examined early
response to determine who does or does not
require radiation post-chemotherapy.
The prognostic importance of early chemotherapy response rather than end of chemotherapy response has led to the use of early response
assessment (after 6–9 weeks) to titrate individual
therapy and dose-dense regimens to maximize
the early response rates. The St. Jude/DFCI/
Stanford Consortium has reported 2-year EFS of
90.8% in early-responding, low-risk patients
with either classical or nodular lymphocytepredominant HL treated with 4 cycles of VAMP
without radiation [51]. The most recent COG
study (AHOD0431) found that early assessment
by PET after one cycle is a predictor of recurrence [41, 52]. The current EuroNet PHL-C1
classical HL trial is evaluating PET activity after
two intensive cycles of OEPA (cumulative dose
of anthracycline is 160 mg/m2) to predict who
does not require radiotherapy [53]. All such
reductions in treatment may increase the risk of
relapse; hence, adverse outcomes such as the
need for high-dose salvage therapy (e.g., stem
cell transplant or high-dose radiation) must be
closely monitored.
G. W. Hall et al.
that eliminated traditional alkylating agents were
not successful and resulted in a 70 and 49% 5-year
EFS for stage III and IV HD, respectively [61].
It has been known for decades that outcome
in HL is optimized by chemotherapeutic dose
intensity. Only recently has this knowledge been
considered a clue to improving outcome [62–
64]. ABVE-PC was developed by the COG (by
adding prednisolone and cyclophosphamide to
ABVE) for the treatment of advanced HL and
dose density was increased by the use of
3-week cycles [17]. This regimen is similar to
dose-dense regimens such as Stanford V and
BEACOPP, developed simultaneously in the
adult groups [18, 19]. BEACOPP and escalated
BEACOPP are dose-intensive regimens with
improved efficacy compared to COPP/
ABVD. Instead of further cumulative dose escalation, the COG and EuroNet PHL take advantage of dose-dense delivery to limit cumulative
cytotoxic therapy. Such dose-intensive regimens
also limit the cumulative dose of agents delivered to the early responders. The GPOH-HD/
EuroNet PHL group has substituted dacarbazine
for procarbazine, resulting in excellent longterm results [40].
ABVE-PC is the backbone for all COG trials.
This dose-dense approach allows for the elimination of procarbazine and the limitation of the doxorubicin and etoposide dose. The first such study
(POG 9425) resulted in 5-year EFS of 84% and
15.1.2.3 Intermediate- and High-Risk
5-year overall survival (OS) of 95% for advanced
(Advanced, Unfavorable)
HL. Early responders (after three cycles of
Disease
ABVE-PC) on this study proceeded directly to
For children with advanced-stage disease, receive 21 Gy regional RT. Others received two
improving efficacy while limiting long-term tox- more cycles (total five ABVE-PC in 15 weeks) prior
icity is even more challenging. The approach in to 21 Gy RT This backbone was used in AHOD0031
pediatric HL has been to increase the number of to evaluate a response-based vs. standard approach
agents so as to limit cumulative doses of indi- to therapy for intermediate-­risk disease and to study
vidual agents. Regimens used in the 1980–1990s augmentation of therapy for high-risk patients with
alternated MOPP/ABVD [29, 54] or used the a slow early response to therapy [65].
hybrid COPP/ABV [34] to avoid the cumulative
Low-dose, involved-site radiation remains a
doses of doxorubicin (300–400 mg/m2) and relevant modality of therapy in high-risk disease.
bleomycin (120–160 mg/m2) associated with six The multicenter trial GPOH-HD95 used OPPA/
to eight cycles of the four-drug ABVD regimen COPP for girls and OEPA/COPP for boys with
[28, 32].
radiation dose determined by end of chemotherMinimalistic dose regimens in combined-­ apy response. For the intermediate- and
modality protocols, such as VEPA (Table 15.4), higher-­risk groups (TG2 and TG4), outcome was
380
AHOD 0031
159
216
67
COG—P9425
(2009)
Chemotherapy
alone
UKCCSG (2002)
68
99
St. Jude’s, Stanford,
DFCI (2004)
CCG 59704
145
77
CS IB/IIB or
Bulky>6 cm
CS III/IV
CS IB, IIB, IIIA2,
IIIB, IV
IIA/IIIA1 “bulk”
CS IV
IIB, IIIB, IV
III, IVB
IIBE, IIIAE, IIIB,
IV
Not I/IIA no bulk,
III, IVB
239
AHOD 0831
IIEB, IIIEA/B,
IIIB, IVA/B
280
382
151
153
Stage
Patients (n)
Group of institution
Combined-modality
trials
Germany, Austria
HD-95 (2001)
(2013)
HD-2002
6–8 ChlVPP
RER: 3 ABVE-PC
SER: 5 ABVE-PC
BEACOPP × 4
RER (F): BEACOPP × 4
RER (M): ABVD × 4
SER: BEACOPP × 4
6 VAMP/COP
SER: Add Iifos/vinor × 2
RER: 5 ABVE-PC
RER: ABVE-PC × 4
SER: ABVE-PC × 4
SER: ABVE-PC × 4 add
DECA × 2
RER:ABVE-PC × 4
2 OEPA, 4 COPDAC
2 OEPA/OPPA
+ 4 COPP
Chemotherapy
Table 15.4 Treatment results for advanced, unfavorable pediatric Hodgkin lymphoma
84
55.2
5
Pediatric Hodgkin Lymphoma
(continued)
[59]
[17]
80.8
5
21 IF
Nonea
86
95
25.5 IF if PR
21 IF
5
93
[57]
[56]
[55]
[40]
[37]
[38]
[34]
[35]
[58]
76
5
3
10
21 GY
21 GY
15 IF if CR
94
80.2
84.3
75.2
79.3
87.9
87
84
Reference
[16]
97
95.9
95.3
97.8
95
97
Follow-up
interval (year)
No
21 GY to residual or
bulky disease
21GY IF RT
No RT
21GY IF RT
21GY IF RT
21GY IF RT
20 ± 10–15 Gy IF RT
PR:20–35, IF CR:
No RT
Radiation (Gy), field
Survival, %
DFS. EFS. or
Overall RFS
15
283
CS IIB, III2A,
IIIB, IV
All except:
IA, IIA non-bulk
IIIB, IVB
81
1712
141
Stage
CS I/IIb, CS IIB,
CS III
CS IV
Patients (n)
394
COPP/ABV, CHOP,
AraC/VP-16
4 MOPP/4 ABVD
Standard arm;
21 IF/4 ABVE-PC
Chemotherapy
6 COPP/ABV
81
79
85
94
96
98
None
None
Randomized: RER: ± RT
SER: ± 2 DECA
Radiation (Gy), field
None
Survival, %
DFS. EFS. or
Overall RFS
100
83
5
4 [60]
3
Follow-up
interval (year)
3
[16]
Reference
[34]
ABVD Adriamycin, bleomycin, vinblastine, and dacarbazine, ABVE-PC Adriamycin, bleomycin, vincristine, etoposide, prednisone, and cyclophosphamide, AEIOP Italian
Association of Hematology and Pediatric Oncology, AraC cytosine arabinoside, CAPTe cyclophosphamide, Adriamycin, prednisone, and teniposide, CCG Children’s Cancer
Group, CCOPP vincristine (Oncovin), procarbazine, and prednisone, ChlVPP chlorambucil, vinblastine, procarbazine, and prednisolone, CHOP cyclophosphamide, hydroxydaunomycin, vincristine (Oncovin), and prednisone, COMP cyclophosphamide, vincristine (Oncovin), methotrexate, and prednisolone, COPP cyclophosphamide, vincristine
(Oncovin), prednisone, and procarbazine, COPP/ABV cyclophosphamide, vincristine (Oncovin), procarbazine, prednisone, Adriamycin, bleomycin, and vinblastine, CR complete response, CS clinical stage, CVPP cyclophosphamide, vinblastine, procarbazine, and prednisone, DFS disease-free survival, EF extended field, EFS event-free survival, HD
Hodgkin’s disease, IF involved field, MDH multicenter trial, MH multicenter Hodgkin’s trial, MOPP nitrogen mustard, vincristine (Oncovin), procarbazine, and prednisone, NR
no response, OEPA vincristine (Oncovin), etoposide, prednisone, and Adriamycin, OPA vincristine (Oncovin), prednisone, and Adriamycin, OPPA vincristine (Oncovin), procarbazine, prednisolone, and Adriamycin, POG Pediatric Oncology Group, PR partial response, PS pathologic stage, R regional, RFS relapse-free survival, RT radiotherapy,
SFOP French Society of Pediatric Oncology, TLI total lymphoid irradiation, UKCCSG United Kingdom Children’s Cancer Study Group, VAMP vinblastine, doxorubicin, methotrexate, and prednisone, VEEP vincristine, etoposide, epirubicin, and prednisolone, VEPA vinblastine, etoposide, prednisone, and Adriamycin, DECA dexamethasone, etoposide,
cisplatin, cytarabine
a
Presence of adverse features = (t)hila, >4 nodal sites, bulk
b
12 patients received 20–35 Gy, IF; 2 received whole lung irradiation
US POG (1997)
Response-based RT
AHOD0031 (2010)
Group of institution
US CCG (2002)
Table 15.4 (continued)
284
G. W. Hall et al.
15
Pediatric Hodgkin Lymphoma
significantly better for those receiving radiation
therapy (TG2, 0.78 vs. 0.92; TG 2 + 3, 0.79 vs.
0.91) [33, 38]. The Children’s Cancer Group
also noted improved outcome for patients treated
with radiation, despite CR at the end of chemotherapy [34, 35]. Kelly et al. [66] reported excellent results using a modified approach to
BEACOPP that reduced doses of chemotherapy
for girls and for boys with a rapid response.
Nonetheless, this regimen is not being used currently because cumulative doses of chemotherapy remain high.
Recent trials in both the COG and in Europe
addressed early-response-directed approaches
to limit the need for radiation. AHOD0031 for
intermediate-­
risk HL used the dose-dense
ABVE-PC regimen to support and evaluate the
concept of an early-response-based algorithm
[60]. This study showed that rapid early response
(RER) could identify a cohort comprising 45%
of patients who did not benefit from radiation.
However, in a subset analysis from this study of
patients with anemia and bulky limited-stage
disease, the EFS was 89.3% for rapid early
responder or complete remission patients who
received IFRT, compared with 77.9% for
patients who did not receive IFRT (P = 0.019)
[67]. For patients who had a slow early response
(SER), a marginal benefit from augmented chemotherapy was observed. The high-risk study
(AHOD-0831) limited radiation fields for rapid
early responders while augmenting therapy for
slow early responders; outcomes were similar to
POG9425 but used less radiation for RER and
less doxorubicin for SER [E].
Adult patients with high risk randomized to
ABVD vs. brentuximab with AVD have been
reported to have a reduced risk of progression,
death, or non-complete response [68], resulting
in approval in the United States for this indication. However, it is not clear that this approach
has an advantage in the setting of pediatric regimens that have had greater efficacy than
ABVC. COG has a randomized, ongoing study
comparing standard ABVE-PC to ABrVE-PC
(Harker-Murray et al.), and the St. Jude
Consortium is similarly evaluating the use of
brentuximab with their backbone therapy.
285
15.1.2.4
Future Considerations
in Classical Pediatric
and Adolescent HL
Progress has been made in the treatment of children with HL with all stages of disease and risk
factors, but several issues remain to be resolved.
Response to chemotherapy may define both the
total amount of chemotherapy required and the
need for radiotherapy (RT). For early-stage
patients, the balance between chemotherapy
dose and radiation exposure continues to be
explored. Restriction of RT to initially involved
lymph nodes (involved-node irradiation or
involved-site irradiation) rather than chains (or
regions) of nodes may affect the balance of risk.
For high-­risk disease, dose-dense chemotherapy
improves efficacy and supports tailoring of therapy to the patient’s response. RT is clearly effective in enhancing the local control of PHL, but
has a dose-dependent toxicity profile favoring a
limited volume/dose approach. Ongoing studies
are needed to assess the role of RT for initial
bulky disease, to residual postchemotherapy
disease (particularly if it is PET negative), and
to involved organs. Carefully designed and
sequential evidence-­based studies are needed to
continue to improve efficacy while limiting
toxicity.
15.1.3 Nodular Lymphocyte-­
Predominant HL (NLPHL)
An indolent, peripheral, NHL-like disease, NLPHL
was recognized in the early 1990s as a clinicopathologically distinct form of HL [69]. Unlike classical HL, NLPHL is a CD20-positive, CD30- and
CD15-negative, B cell lymphoma that is not associated with EBV genomic integration. There is a
distinct male predominance (ratio 2–3:1) with
nearly 90% of pediatric patients having early-stage
disease (IA/IIA). A higher percentage (10–20%) of
children have NLPHL [3] compared to adults
(3–8%) [70], and although >50% of pediatric and
adolescent cases are under the age of 14 years [71],
the incidence peaks between 14 and 18 years.
Peripheral lymphadenopathy is the most common
presentation involving the axilla, cervical, and
G. W. Hall et al.
286
inguinal regions, often present for months or years.
Rarely is advanced or central disease seen.
Adults with early-stage NLPHL are treated
with involved-field radiotherapy, standard cHL
therapy, or combined-modality therapy. Children
have, until 2005 and the start of NLPHL-specific
clinical trials, received standard pediatric cHL
therapy with combined-modality chemoradiotherapy [72], which is excessively toxic.
Morbidity, even mortality, secondary to
repeated courses of intensive therapy to eradicate
this indolent, usually nonfatal disease has resulted
in a drive to reduce the intensity of therapy to
avoid late effects [71].
Children with fully resected early-stage
nLPHD have been cured without the need for any
chemoradiotherapy [73–76], but the specific situations in which this strategy is appropriate are
currently under investigation. Two nonrandomized clinical trials, EuroNetPHL-LP1 and COG’s
AHOD03P1, have looked at reducing the toxicity
of upfront therapy for early-stage disease (stage I
and II) [73, 74]. As salvage therapy is effective for
late or even multiple relapses which generally
recur at the original site of disease with no stage
upgrade, OS is expected to remain near to 100%
[77]. The EuroNetPHL-LP1 used surgical resection alone or low-dose anthracycline-free CVP
chemotherapy for non-resectable disease, and
COG’s AHOD03P1 used AVPC (equivalent to
CHOP) with selective radiotherapy. Excellent
EFS rates of 60–75% with no or low-dose chemotherapy have been obtained and only 10% of COG
patients received RT, maintaining 100% OS [78].
Because of transformation rates of approximately 5% to aggressive B-NHL [79] in adults,
usually diffuse large B cell lymphoma [80], concerns regarding reduced therapy that could potentially allow persistence of the CD20 clone and
increased transformation rates remain. In theory,
the addition of rituximab would help to specifically eradicate the CD20 clone and reduce transformation rates. However, transformation rates in
children are not known but appear extremely low.
Rituximab has been studied in adults for use in
this and all other CD20-positive lymphomas [81].
The pediatric community have traditionally been
wary about using rituximab in young children
because of impact on immune status/memory. As
early-stage NLPHL is viewed as a highly curable disease with minimal chemotherapy or surgery alone,
the use of rituximab has been reserved for treating
more aggressive, advanced, or relapsed disease.
Assessing the impact of adjuvant rituximab therapy
on EFS and transformation rates in children within a
randomized clinical trial has been the unattainable
aim of clinicians for well over a decade. The reluctance of the pediatric community to use rituximab in
this and other CD20+ lymphomas is abating.
Current proposed clinical trials using low-­
dose NHL-like therapy including anti-CD20
therapy are focused on the natural history, establishing risk categories, variant histologies, and
transformation rates, with biological substudies
looking at specific molecular characteristics.
15.1.4 Recurrence, Relapse,
and Salvage in PHL
15.1.4.1 Introduction
Relapsed and refractory classical Hodgkin lymphoma (HL) remains a clinical and therapeutic
challenge. Approximately 10% of patients with
early-stage and up to 30% with advanced-stage
disease relapse after first-line chemotherapy.
Cure can still be achieved in a substantial proportion of patients with recurrent disease, but
there is no uniform approach to salvage therapy.
The optimal salvage treatment has not been
defined in children and adolescents as there are no
randomized trials defining the “best” salvage chemotherapy regimen or comparing standard-­dose
chemotherapy (SDCT) vs. high-dose chemotherapy and autologous stem cell transplant (HDCT/
ASCT), which is often considered the standard of
care in adult practice. Pediatric practice adopts a
more individualized risk-stratified and responseadapted approach to salvage treatment with both
non-transplant (SDCT plus radiotherapy) and
transplant (SDCT plus HDCT/ASCT) salvage.
At the point of relapse, a full disease reassessment including histologic confirmation is mandatory and then an analysis of pre-salvage risk
factors is undertaken. All patients have a common starting point with re-induction SDCT and
this is followed by consolidation treatment. The
choice of consolidation treatment is guided by
15
Pediatric Hodgkin Lymphoma
risk stratification based on prognostic factors as
well as an assessment of chemosensitivity which
is commonly done after two cycles of SDCT and
includes FDG-PET response. Achieving a complete metabolic remission on FDG-PET prior to
consolidation has been shown to be highly prognostic in the relapse setting and is considered to
be a major goal of re-induction SDCT [82].
Consolidation after SDCT will be radiotherapy
only in “low-risk” relapse or HDCT/ASCT in
“standard-risk” relapse, and these two strategies
will be appropriate for the vast majority of
relapse/progressive HL. A small number of
patients are refractory to SDCT and do not
achieve a CR with two or more lines of SDCT
and these are “high-risk” patients [82].
Consolidation in these high-risk patients may be
either conventional HDCT/ASCT possibly with
post-HDCT consolidation RT or maintenance-­
targeted therapy such as brentuximab vedotin, or
alternative experimental approaches may be
applied including novel agents such as checkpoint inhibitors or allogeneic transplantation.
15.1.4.2
Standard-Dose Salvage
Chemotherapy Regimens
After recurrence is noted, the first step is reinduction with a SDCT salvage regimen. There is
no “best” chemotherapy regimen at salvage,
and there are no randomized studies comparing standard-­dose chemotherapy regimens. The
choice of regimen should take account of primary therapy, use of non-cross-resistant drugs,
and cumulative drug toxicities. The aim of salvage therapy is to obtain cytoreduction and to
demonstrate chemosensitivity which is done
most accurately now with FDG-PET as firstline treatment. It also facilitates collection of
peripheral stem cells for ASCT. Salvage
regimes can be divided into intensive conventional regimens1 ­(mini-­BEAM), cisplatinbased regimens2 (ESHAP, DHAP [ESHAP,
DHAP, APPE, DECAL]), ifosfamide-­
based
Mini-BEAM; BCNU, etoposide, cytarabine, melphalan
ESHAP, etoposide, methylprednisolone, cytarabine, cisplatin; DHAP, dexamethasone, cytarabine, cisplatin;
APPE, cytarabine, cisplatin, prednisone, etoposide;
DECAL, cytarabine, cisplatin, prednisone, etoposide,
asparaginase
1
2
287
regimens3(EPIC, IEP, ICE, IV), or others4 (GV,
IGEV). The COG uses IV as its standard regimen because of efficacy and with the intent of
avoiding etoposide-induced secondary malignancy after stem cell transplantation [83]. In
Europe, alternating IEP/ABVD was used in the
EuroNet-PHL-R1 trial but more recently the
IGEV regimen has been widely adopted. The
decision to continue salvage therapy with RT
consolidation vs. HDCT/ASCT is based on
assessment of predictive factors.
15.1.4.3
Prognostic Factors
at Relapse in Pediatric HL:
Standard-­Dose
Chemoradiotherapy Vs.
High-Dose Chemotherapy/
Stem Cell Transplantation
Prognostic factors at relapse may be used to allocate patients to a risk-stratified salvage approach.
This is in contrast to adult practice where consolidation with HDCT/ASCT is considered standard
of care. There are currently no universally
accepted prognostic criteria in children (or adults)
defining individualized salvage treatment plans.
Factors which are prognostically important
include time to relapse, prior treatment in first
line, stage/disease burden at relapse, and response
to salvage chemotherapy. In children, low-risk
patients may be salvaged with RT consolidation
only, while standard-risk patients are salvaged
with HDCT. The cut point between low- and standard-risk patients is not universally defined. In
Europe, low-risk patients salvaged with SDCT
plus RT only include those with early relapse after
up to 4 cycles of chemotherapy and late relapse
after up to 6 cycles with all of the following: nodal
relapse, no prior RT (or relapse only outside prior
RT fields), consolidation RT that has acceptable
toxicity (i.e., no excessive RT fields), and chemotherapy-responsive disease. All other patients
have intensification with HDCT/ASCT.
EPIC, etoposide, vincristine epirubicin, prednisolone;
IEP, ifosfamide, etoposide, prednisolone; ICE, ifosfamide, carboplatin, etoposide; IV, ifosfamide,
vinorelbine
4
GV gemcitabine, vinorelbine; IGEV, ifosfamide, gemcitabine, vinorelbine, prednisolone
3
G. W. Hall et al.
288
Time to relapse from end of first-line treatment is the most important pretreatment risk factor and highly significant for OS and EFS in
pediatric studies [84–86] and dominated all other
prognostic factors in multivariate analysis of the
ST-HD-86 trial, the largest prospective pediatric
relapse trial published to date [87], with DFS of
41, 55, and 86% for those with refractory disease,
early relapse, and late relapse, respectively. This
study showed that salvage can be risk adapted
because subgroups with markedly better or worse
prognosis can be defined. Stage IV and extranodal disease were also associated with lower OS.
A recent French experience [88] found the
only relevant prognostic factors to be time to
relapse and chemoresistance with primary progressive HL having an EFS <40% compared with
approximately 80% in late relapse and chemosensitivity (CR or PR >70%) to salvage associated with a DFS of 77% vs. 10% with poor
response (p < 0.0001). Chemosensitivity to
SDCT and disease status at transplantation are
also predictive of outcome. In one study, 5-year
FFS was 35% for patients with chemosensitive
disease vs. 9% with chemoresistant disease [84].
Another group found 68% OS and 59% FFS at
5 years in chemosensitive patients vs. 18% and
0% in chemoresistant patients [85]. Several particularly adverse factors have been noted.
Chemoresistant patients had 5-year FFS of 0%
with HDCT/ASCT [85]. Adolescents with B
symptoms at recurrence had poor OS even after
HDCT/ASCT (11-year OS 27% with B symptoms vs. 60% without) [89]. No difference in OS
or FFS between age subgroups or in comparison
with adult cohorts has been reported by several
studies [84, 85, 90]. Of note, many of these studies did not incorporate FDG-PET response
assessment which is now well recognized as the
most important prognostic factor, which may
overcome the significance of some factors as is
the case in first-line treatment [91].
15.1.4.4
Role of Radiotherapy
in Relapsed Hodgkin
Lymphoma
Radiotherapy has an important role in salvage,
but must be individualized based on previous
radiation exposure, in or out of field recurrence,
stage at recurrence, and the toxicities of total treatment burden [92]. Increasing numbers of patients
are RT naïve at relapse as the use of RT is increasingly restricted in first-line treatment and RT fields
are also becoming highly restricted in some firstline trials to FDG-PET-positive residua. Therefore,
at relapse many patients have never received RT,
and some other patients may relapse in prior disease sites that have never received RT because they
received focal targeted RT only. Salvage with RT
alone is generally not recommended, but integration of RT in salvage is relevant in two contexts:
1. As consolidation treatment in low-risk group
patients after SDCT.
2. In selected patients as consolidation after
HDCT/ASCT
15.1.4.5
High-Dose Chemotherapy
and Autologous Stem Cell
Transplant
COG protocols have studied HDCT/ASCT and
immunomodulatory therapy in all patients
except the lowest-risk group (late relapse without bulky disease or B symptom in those initially treated for IA/IIA disease with minimal
systemic therapy) [93]. In Europe, HDCT/
ASCT has a recognized role in salvage for those
with higher-risk features, namely, all primary
progressive HL and early relapse after 6 cycles
of first-line chemotherapy, all relapse with poor
response to reinduction, and finally those
patients in whom RT consolidation is either not
feasible (advanced-­stage relapse) or too toxic
(extensive RT fields required or re-irradiation of
prior irradiated sites). Patients without high-risk
features and who achieve a complete FDG-PETdefined response after two cycles of SDCT may
receive only consolidation SDCT plus RT.
There are no studies that define the most effective HDCT. BEAM and CVB (cyclophosphamide, etoposide, carmustine) are commonly
used. TBI-containing regimens confer no benefit
and are associated with increased toxicity and
late effects. Transplant-related mortality is down
to 0–2% in some series. A higher TRM rate has
been associated with history of atopy, thoracic
irradiation, multiple chemotherapy regimens, and
multiple relapses.
15
Pediatric Hodgkin Lymphoma
Series with HDCT/ASCT in pediatric and
adolescent patients are small and report EFS
rates of 31–67% [84, 85, 90, 94]; outcome for
children is similar to adults with HDCT/ASCT
[84, 90]. Studies that evaluate survival benefit
rather than event-free survival after disease
recurrence often rely on transplant after second
or later recurrence to achieve good OS [85, 95].
Patients with primary progressive disease and
those resistant to salvage regimens remain a
huge challenge. SDCT with radiotherapy will
not afford a chance of cure, but even HDCT/
ASCT is inadequate therapy for most such
patients. New approaches to such patients, such
as use of post-HDCT consolidation maintenance-targeted treatment, were tested in the
Aethera trial with up to 16 cycles of brentuximab
vedotin or post-HDCT radiotherapy which is
also an option to minimize further relapse.
Allogeneic SCT or immunomodulatory therapy
may prove beneficial [93].
Long-term follow-up is required post-HDCT
for detection of late relapse and development of
second cancers, which have been reported at a
rate of 5–10% at 5 years and substantially higher
at 20 years or more in some series. Thirty-eight
percent of deaths occurred 4–12 years after
ASCT; 85% of relapses occur within 2 years of
ASCT [86].
15.1.4.6
High-Dose Chemotherapy
and Allogeneic Stem Cell
Transplantation
The role of allogeneic transplant in relapsed HL
remains unknown. The poor outcome with
HDCT/ASCT in chemotherapy poor responders
to salvage and those who remain FDG-PET
­positive after salvage has resulted in exploration
of alloSCT. Allogeneic transplantation is not recommended as the initial transplant approach outside of a clinical trial setting [96] due to the high
non-relapse mortality (NRM) rate, mainly caused
by graft vs. host disease and infection. Reduced
intensity conditioning (RIC) ameliorates the
NRM while maintaining theoretical graft vs.
lymphoma effect. Allogeneic-SCT may be an
option for relapse post-HDCT/ASCT and for
patients with refractory advanced-stage HL and
chemoresistant disease at salvage.
289
Children and adolescents allografted for HL
had an OS of 45% and PFS of 30% at 5 years
[97]. All were heavily pretreated, almost half
with HDCT/ASCT. Those with chemosensitive
disease and good performance status achieved
3-year OS of 83% and PFS of 60%. NRM was
21 ± 4% in both the RIC and myeloablative conditioning groups. RIC was associated with a significantly higher relapse risk compared to
myeloablative conditioning. Graft vs. host disease did not affect relapse rate.
Although studies based on registry data are
useful, prospective trials are required to gain a
better understanding of the role of allogeneic
transplantation. The indications, optimal time
point, conditioning regimen, and GVHD prophylaxis still need to be better defined. With the
advent of newer immunotherapy agents, including checkpoint inhibitors, the role of alloSCT
globally in HL is under review and the numbers
of such transplants are declining globally.
15.1.4.7
Brentuximab Vedotin
and Checkpoint Inhibitors
In recent years there have been two early-phase
pediatric trials investigating novel agents in children. The first is the phase I/II pediatric trial
(ClinicalTrials.gov number NCT01492088)
investigating single-agent brentuximab vedotin
in R/R HL and anaplastic large cell lymphoma
[98]. The recommended phase II dose was
1.8 mg/kg as in adults and the ORR was 47%
(CR rate 33%, PR rate 12%) in HL patients and
toxicity was manageable. This compares with the
pivotal phase II study in adults where the ORR
was 75% (CR rate 34%) [99]. The second is the
ongoing risk-stratified and response-adapted
phase II salvage trial in first R/R HL of nivolumab
plus brentuximab vedotin followed by bendamustine plus brentuximab vedotin in poor initial
responders in first R/R HL in children and young
adults (Checkmate 744 trial, AHOD1721;
NCT02927769) [100]. The preliminary results of
this study are recently presented showing 64% of
patients achieved a CMR after brentuximab
vedotin plus nivolumab. Of those inadequate
responders that switched to second-line brentuximab vedotin plus bendamustine, all achieved a
CMR after 2 cycles of this intensification.
290
The overall CMR rate with either first or second
salvage in this trial was 86%, demonstrating that
only a small number of patients cannot achieve a
CMR pre-HDCT with these combinations.
Treatments that block the interaction between
programmed death-1 (PD-1) and its ligands have
shown high levels of activity in adults with
HL. The anti-PD-1 antibody nivolumab induced
objective responses in 20 of 23 adult patients
(87%) with relapsed HL [101]. Another anti-­
PD-­
1 antibody, pembrolizumab, produced an
objective response rate of 65% in 31 heavily pretreated adult patients with Hodgkin lymphoma
who relapsed after receiving brentuximab vedotin [102]. These agents may be used as a bridge to
transplant, as post-HDCT maintenance brentuximab vedotin, or as alternatives to conventional
SDCT. These novel agents when used as a single
agent achieve CR rates of 19–33%, but in combination achieve higher CR rates as in the
Checkmate trial [100]. An interesting combination is brentuximab vedotin plus bendamustine
[103] which achieves CR rates in excess of 75%
which means that most patients can achieve a CR
prior to HDCT making the use of alloSCT which
is often used in patients that cannot achieve a CR
less appealing.
15.1.5 Late Effects
Long-term adverse sequelae of greatest concern in
children treated for HL (particularly with regimens
including high-dose radiation) include impairment
of muscle and bone development [5] and injury to
the lungs [104], heart [105], thyroid gland [11, 12],
and reproductive organs [106]. Cardiovascular dysfunction, pulmonary fibrosis, and secondary malignancies significantly compromise the quality and
length of life in survivors [107].
15.1.5.1 Cardiac Toxicities
High-dose (>30 Gy) radiation to the mediastinum
has been associated with significant long-term
effects in patients with HL. Stanford investigators reported that the actuarial risk of developing
cardiac disease necessitating pericardiectomy
was 4% at 17 years in a series of long-term
G. W. Hall et al.
survivors of childhood HL who had received
high-­dose radiation [14]. Screening echocardiogram, exercise stress test, and resting and 24-h
ECG identified numerous clinically significant
cardiac abnormalities in HL patients who had
mediastinal irradiation at a median age of
16.5 years (range, 6.4–25 years). Significant valvular defects were detected in 42%, autonomic
dysfunction in 57%, persistent tachycardia in
31%, and reduced hemodynamic response to
exercise in 27% of patients [108]. With the introduction of techniques that reduce the radiation
dosage to the heart, rates of radiation-associated
cardiac injury have declined dramatically.
Mediastinal irradiation given for HL may further predispose patients with PHL to
anthracycline-­related myocardiopathy [14, 109].
Cardiac dysfunction after anthracycline therapy
itself is notable, with the highest risk in those
receiving high cumulative doses or in young children who may be affected by an adverse effect on
cardiac myocyte growth [14, 109]. Fortunately,
most pHL patients are adolescents and current
pHL regimens doses are significantly lower than
those used in adult ABVD regimens.
15.1.5.2 Pulmonary Toxicities
Chronic pneumonitis and pulmonary fibrosis
should be rare in the current era of treatment for
primary HL (Fig. 15.1). Predisposing therapies
include thoracic radiation and bleomycin chemotherapy [104, 105]. The bleomycin in ABVD
can cause both acute pulmonary compromise
and late pulmonary fibrosis and can be augmented by the fibrosis that can be associated
with pulmonary radiation. Asymptomatic pulmonary dysfunction that improves over time has
been observed after contemporary combinedmodality treatment.
15.1.5.3 Thyroid Toxicities
Thyroid sequelae are common after RT for
PHL. Hypothyroidism, hyperthyroidism, thyroid
nodules, and thyroid cancer have been observed in
long-term survivors [11, 12]. Of these, hypothyroidism, particularly compensated hypothyroidism, defined as thyroid-stimulating hormone (TSH)
elevation in the presence of a normal thyroxine
15
Pediatric Hodgkin Lymphoma
291
(T4) level, is the most common thyroid abnormality. The primary risk factor for hypothyroidism is
higher cumulative radiation dosage; the influence
of age remains controversial [11, 12]. As many as
78% of patients treated with radiation dosages
greater than 26 Gy demonstrate thyroid dysfunction, as indicated by elevated TSH levels [11].
15.1.5.4 Secondary Malignancies
The overall cumulative risk of developing a
subsequent malignancy after treatment for
PHL has been reported to range from 7% to
10% at 15 years from diagnosis and rises to
16–28% by 20 years (Table 15.5) [116]; these
data are based on patients treated in earlier
decades. The most common secondary malignancies historically included both secondary
acute myeloid leukemia (MDS/secondary
AML) and solid tumors. However, leukemias
are now infrequent due to changes in chemotherapy. Female breast cancer is a particular
concern but is likely to be less common with
current radiation doses and techniques, since it
is associated with RT fields that include breast
tissue (especially mantle fields) and higher
radiation doses (Fig. 15.1).
15.1.6 Summary/Future Directions
Tremendous strides have been made in treating
children with HL, both in terms of cure and
reduction of toxicity. Devising new strategies to
treat children with HL is problematic because of
the overall success of current treatment regimens.
However, grouping patients into different risk
categories, using response-based therapy and
newer imaging techniques, allows investigators
to construct protocols intended to diminish
therapy-­induced toxicity for patients with favorable prognoses. These protocols also aim to
improve efficacy of treatment for patients with
intermediate and unfavorable prognoses.
Unfortunately, the ability to conduct clinical trials, where the difference in survival between
treatment arms is likely to be small, is compromised by the large patient numbers required to
detect such differences. If a reduction in treatment toxicity is the intended goal of a new regimen, then many years of follow-up are necessary
to prove efficacy. For patients with refractory, or
multiple relapsed disease, phase II studies investigating the use of monoclonal anti-CD30 and
anti-PD-1 antibodies alone and in combination,
and with other checkpoint inhibitors, in children
and adolescents are ongoing internationally. The
importance of investigators working together
throughout the world to share data and new treatment approaches in order to cure children with
HL safely is clear.
Acknowledgments Thanks to Laura Finger, Rochester,
for her help with the final draft and references of the third
edition 2019.
Table 15.5 Secondary cancers after childhood HL
Reference
Stanford [110]
Cohort
size
694
LESG [111]
1641
[112]
Roswell [113]
LESG [114]
US/European [115]
University of Rochester/Johns
Hopkins/University of Florida/
St. Jude/Dana-Farber [116]
Number of
Time period secondary
cancers
studied
1960–1995
59
1136
182
1380
5925
1940s to
1991
1955–1986
1960–1989
1955–1986
1935–1994
62
162
28
135
195
930
1960–1990
102
Cumulative incidence
(%) (years)
Males, 9.7%
(20 years); females,
16.8% (20 years)
18% (30 years)
26.4% (40 years)
26.7% (30 years)
31.2% (30 years)
Solid tumors: 11.7%
(25 years)
19% (25 years)
Standardized
incidence ratio
Males, 10.6;
females, 15.4
7.7
9.4
17.9
7.7
Males, 8.41;
females, 19.93
G. W. Hall et al.
292
References
1. Rubin P, Williams JP, Devesa SS, Travis LB,
Constine LS (2010) Cancer genesis across the age
spectrum: associations with tissue development,
maintenance, and senescence. Semin Radiat Oncol
20:3–11
2. Punnett A, Tsang RW, Hodgson DC (2010) Hodgkin
lymphoma across the age spectrum: epidemiology, therapy, and late effects. Semin Radiat Oncol
20:30–44
3. Hochberg J, Waxman IM, Kelly KM, Morris E,
Cairo MS (2009) Adolescent non-Hodgkin lymphoma and Hodgkin lymphoma: state of the science.
Br J Haematol 144:24–40
4. Foltz LM, Song KW, Connors JM (2006)
Hodgkin’s lymphoma in adolescents. J Clin Oncol
24:2520–2526
5. Donaldson SS,
Download